Synthesis Characterization and Applications of Colloidal Supported

University of Rhode Island

DigitalCommons@URI

Open Access Dissertations

2013

Synthesis Characterization and Applications of

Colloidal Supported Metal Nanoparticles

Kalyani Gude

University of Rhode Island , kalyanigude@gmail.com

Follow this and additional works at: http://digitalcommons.uri.edu/oa_diss

Terms of Use

All rights reserved under copyright.

Recommended Citation

Gude, Kalyani, "Synthesis Characterization and Applications of Colloidal Supported Metal Nanoparticles" (2013).

Open Access

Dissertations.

Paper 55.

This Dissertation is brought to you for free and open access by DigitalCommons@URI. It has been accepted for inclusion in Open Access Dissertations by an authorized administrator of DigitalCommons@URI. For more information, please contact digitalcommons@etal.uri.edu

.

SYNTHESIS CHARACTERIZATION AND APPLICATIONS OF COLLOIDAL

SUPPORTED METAL NANOPARTICLES

BY

KALYANI GUDE

A DISSERTATION SUBMITTED IN PARTIAL FULFILLMENT OF THE

REQUIREMENTS FOR DOCTOR OF PHILOSOPHY

IN

CHEMISTRY

UNIVERSITY OF RHODE ISLAND

2013

DOCTOR OF PHILOSOPHY IN CHEMISTRY

OF

KALYANI GUDE

APPROVED

Thesis Committee:

Major Professor Radha Narayanan

Louis J. Kirschenbaum

Vinka Craver

Nasser H. Zawia

DEAN OF THE GRADUATE SCHOOL

UNIVERSITY OF RHODE ISLAND

2013

ABSTRACT

Nanocatalysts show excellent catalytic activity because of their immense surface-to-volume ratio. In homogeneous nanocatalysis, the catalyst remains in the same phase as the reactants and products. Traditionally available homogeneous nanocatalysts include colloidal suspensions of transition metal nanoparticles. In heterogeneous nanocatalysis, the catalyst and the reactants are in different phases.

Traditionally available heterogeneous nanocatalysts include metal nanoparticles adsorbed onto bulk supports such as silica, alumina, carbon etc. The main advantage associated with homogeneous nanocatalysts is that they have high selectivities compared to heterogeneous nanocatalysts. The main disadvantages associated with homogeneous nanocatalysts include poor thermal stability, metal contamination and difficulty recovering the catalyst. The heterogeneous nanocatalysts on the other hand have good thermal stabilities and good catalyst recoveries. We focused on synthesizing a new type of catalyst that we termed as colloidal supported metal nanoparticles (CSMNs). The CSMNs act as an intermediate class of nanocatalysts that have properties associated with both homogeneous and heterogeneous nanocatalysts.

The synthesized CSMNs are characterized using TEM and EDS. The CSMNs are used as nanocatalysts for the Suzuki and Heck cross-coupling reactions. The kinetics of these reactions is monitored using high performance liquid chromatography (HPLC).

The catalytic activity, stability and the recycling potential of the CSMNs is investigated in our study.

ACKNOWLEDGMENTS

I wish to take this opportunity to thank the people who have been instrumental in making this dissertation possible. I thank my research advisor Dr. Radha Narayanan for being supportive and approachable from the day I joined her lab group. I am thankful to her for mentoring me and making me feel passionate about the science of nanoparticles.

I thank my committee members Dr. Vinka Craver, Dr. Louis Kirschenbaum,

Dr. David Worthen and Dr. Sze Yang for the time, invaluable inputs and guidance on aspects related to research, academics and career. I am very fortunate to have the guidance from all my committee members and I truly and honestly appreciate it.

I specially thank my previous advisor Professor Jonathan Blitz for instilling in me the passion towards research and for the encouragement given.

I thank the faculty and staff at the Department of Chemistry for helping me throughout the time I am here. I specially thank Tim Wasco for helping me with the instrumental analysis laboratory.

I thank my past and present group members particularly, Allison, Ben and Nik for their willingness to help me around the lab. I also thank my good friends Liu,

Mengyun, Sravanthi and many others for their support and help.

I am grateful to my Mom, Dad, sisters and in-laws for their endless love and support. I specially thank Dinesh for standing by me, believing me and more importantly patiently listening to the problems I encountered. Without his encouragement and emotional support, I wouldn’t have been able to complete this.

Thank you Dinesh for motivating me and giving me the confidence that I can do it! iii

PREFACE

This dissertation is written utilizing a manuscript format. The first chapter is an introduction about nanocatalysis. The second chapter entitled “Synthesis and

Characterization of Colloidal Supported Metal Nanoparticlees (CSMNs) as Potential

Intermediate Nanocatalysts” was published in Journal of Physical Chemistry C.

( Journal of Physical Chemistry C , 2010 , 114 , 6256-6362) in March 2010. The third chapter entitled “Colloidal Supported Metal Nanoparticles (CSMNs) as Effective

Nanocatalysts for Liquid-Phase Suzuki Cross-Coupling Reactions” was published in the Journal of Physical Chemistry C. ( Journal of Physical Chemistry C , 2011 , 115 ,

12716-12725) in June 2011. The fourth chapter entitled “Functionalized Colloidal

Supported Metal Nanoparticles (CSMNs) as efficient Recyclable Nanocatalysts for

Liquid-Phase Cross-Coupling Reactions” is submitted for publication in Journal of

Physical Chemistry C. iv

TABLE OF CONTENTS

ABSTRACT .................................................................................................................. ii

ACKNOWLEDGMENTS…………………………………………………………….iii

PREFACE ...................................................................................................................... iv

TABLE OF CONTENTS ............................................................................................... v

LIST OF FIGURES .................................................................................................... viii

Chapter 1. Introduction to Nanocatalysis ....................................................................... 1

Synthesis of Colloidal Metal Nanoparticles…..........................................................2

Chemical Reduction……………………………………………………………3

Thermal and Photochemical Decomposition......................................................5

Electrochemical Reduction………………………………………..…………...6

Stabilization of Metal Nanoparticles… .................................................................... 7

Electrostatic Stabilization...................................................................................7

Steric Stabilization..............................................................................................8

Electrosteric Stabilization……………...............................................................9

Stabilization by Ligands……………………………………………………......9

Stabilization using Dendrimers……………………………………………….10

Supported Nanoparticle Catalysts……………………………………………...…11

Supports………………………………………………………………..………....13

Nanocatalysts for Suzuki and Heck Cross-Coupling Reactions…………………14

Suzuki Reaction…………………………………………………………....…14

Heck Reaction………………………………………………………………...16

Homogeneous and Heterogeneous Catalysis…………………………………….17 v

References……………………………………………………………………….20

Chapter 2. Synthesis and Characterization of Colloidal Supported Metal Nanoparticles

(CSMNs) as Potential Intermediate Nanocatalysts…………………………………...29

Abstract…………………………………………………………………………..29

Introduction……………………………………………………………………....30

Experimental……………………………………………………………………..31

Results and Discussion…………………………………………….……………..34

Conclusions……………………………………………………………………....41

References………………………………………………………………….…….49

Chapter 3.

Colloidal Supported Metal Nanoparticles (CSMNs) as Effective

Nanocatalysts for Liquid-Phase Suzuki Cross-Coupling Reactions………………….54

Abstract…………………………………………………………………………..54

Introduction.……………………………………………………………………..55

Experimental……………………………………………………………………..58

Results and Discussion…………………………………………………………..62

Conclusion……………………………………………………………………….73

References……………………………………………………………………….86

Chapter 4. Functionalized Colloidal Supported Metal Nanoparticles (CSMNs) as

Efficient Recyclable Nanocatalysts for Liquid-Phase Heck Reaction……………..…93

Abstract…………………………………………………………………………..93

Introduction.……………………………………………………………………..94

Experimental……………………………………………………………………..97 vi

Results and Discussion………………………………………………………....101

Conclusion…………………………………………………………………...…108

References……………………………………………………………………...116

Bibilography………………………………………………………………………...119 vii

LIST OF FIGURES

Chapter 2

Scheme 1 Four step process for the synthesis of colloidal supported metal nanoparticles (CSMNs)……………………………………………….………43

Figure 2-1. TEM image of the silica colloids synthesized with the Stoeber method (a) and the size distribution histogram (b)………………………...…44

Figure 2-2. TEM image of the palladium nanoparticles (a) and the size distribution histogram (b)……………………………………………………..44

Figure 2-3. TEM images and EDS spectrum of CSMNs prepared with silica colloids functionalized with 3-mercaptopropyl trimethoxysilane and palladium nanoparticles…...……………………………………………………………..45

Figure 2-4. TEM image of CSMNs prepared using APTES functionalized silica colloids (a) TEM image of the CSMNs at a higher magnification

(b)……………………………………………………………………………..46

Figure 2-5. TEM image of the silica colloids before and after functionalization with 3-mercaptopropyl trimethoxysilane……………………………………..46 viii

Figure 2-6. TEM image of the silica colloids before and after functionalization with 3-aminopropyl triethoxysilane………………………………………......47

Figure 2-7. Kinetics of the CSMNs prepared with palladium nanoparticles

(NPs) bound to silica colloids (SC) functionalized with the amine linker……47

Figure 2-8. TEM images (a-b) and EDS spectrum (c) of the CSMNs after the

Suzuki reaction………………………………………………………………..48

Chapter 3

Figure 3-1. Example TEM image showing the palladium nanoparticles covalently bound to amine functionalized silica colloid suspensions (a) and

EDS spectrum which reflects the presence of both Pd and Si (b)…………….76

Figure 3-2. Examples of TEM images showing palladium nanoparticles directly adsorbed onto the surface of the silica colloid suspensions (a) and EDS spectrum, which reflects the presence of both Pd and Si (b)…………………77

Figure 3-3. A representative TEM image showing the palladium nanoparticles covalently bound to APTES functionalized dried silica colloid suspensions (a) and EDS spectrum, which reflects the presence of both Pd and Si (b)……….78 ix

Figure 3-4. A representative TEM image showing the palladium nanoparticles adsorbed onto the unfunctionalized dried silica colloid suspensions (a) and

EDS spectrum which reflects the presence of both Pd and Si (b)…………….79

Figure 3-5. Example of TEM image showing APTES functionalized bulk silica with palladium nanoparticles covalently attached to its surface (a) and EDS spectrum which reflects the presence of both Pd and Si (c)…………………80

Figure 3-6. Example of TEM image showing unfunctionalized bulk silica with palladium nanoparticles adsorbed to its surface (a) and EDS spectrum which reflects the presence of both Pd and Si (b)…………………...……………….81

Figure 3-7. Comparisons of the catalytic activity with wet APTES functionalized Si colloids reacted with Pd nanoparticles that are air-dried, dry

APTES functionalized Si colloids resuspended in doubly deionized water and reacted with Pd nanoparticles that are air-dried, and bulk dry APTES functionalized Si dispersed in doubly deionized water and reacted with Pd nanoparticles that are air-dried………………………………………………..82

Figure 3-8. Comparisons of the catalytic activity with wet unfunctionalized Si colloids reacted with Pd NPs that are air-dried, dry unfunctionalized Si colloids resuspended in doubly deionized water and reacted with Pd NPs that are airx

dried, and unfunctionalized bulk dry Si dispersed in doubly deionized water and reacted with Pd NPs that are air-dried……………………………………82

Figure 3-9. Comparisons of the catalytic activity with wet Si colloids reacted with Pd NPs without linker that are air-dried and palladium nanoparticles in colloidal solution…………………………………………………………...…83

Figure 3-10. Comparisons of the catalytic activity with wet unfunctionalized Si colloids reacted with Pd nanoparticles that are air-dried, and two different wet

APTES functionalized Si colloids reacted with Pd nanoparticles that are airdried…………………………………………………………………………..83

Figure 3-11. Comparisons of the catalytic activity with unfunctionalized dried

Si colloids resuspended in doubly deionized water that are air-dried and

APTES functionalized dried Si colloids resuspended in doubly deionized water that are air-dried……………………………………………………………....84

Figure 3-12. Comparisons of the catalytic activity with unfunctionalized bulk dry Si dispersed in doubly deionized water that are air-dried and APTES functionalized bulk dry Si dispersed in doubly deionized water that are airdried…………………………………………………………………………84 xi

Figure 3-13. Comparison of the catalytic activity of the supernatant solution after refluxing the CSMNs in 3:1 acetonitrile and centrifuging to that of the

CSMNs used as catalysts for the Suzuki reaction…………………………….85

Chapter 4

Figure 4-1. TEM image of the palladium nanoparticles (1a) Size distribution of

palladium nanoparticles (1b)………………………………………………...110

Figure 4-2. TEM image showing the palladium nanoparticles supported onto

unfunctionalized silica colloid suspensions (a) Size distribution of palladium

nanoparticles supported on unfunctionalised silica colloids (b)…………….110

Figure 4-3.

TEM image showing the palladium nanoparticles supported onto unfunctionalized silica colloid suspensions after the first cycle of heck reaction

(a) Size distribution of palladium nanoparticles supported on unfunctionalised

silica colloids after the first cycle of heck reaction (b)……………………..111

Figure 4-4.

TEM image showing the palladium nanoparticles supported onto unfunctionalized silica colloid suspensions after the second cycle of heck reaction (a) TEM image showing the palladium nanoparticles supported onto unfunctionalized silica colloid suspensions after the third cycle of heck reaction (b)…………………………………………………………………..111

Figure 4-5.

TEM image showing the palladium nanoparticles supported onto

APTES functionalized silica colloid suspensions (a) Size distribution of xii

palladium nanoparticles supported onto APTES functionalized silica colloids

(b)……………………………………………………………………………112

Figure 4-6. TEM image showing the palladium nanoparticles supported onto

APTES functionalized silica colloid suspensions after the first cycle of heck reaction (a) Size distribution of palladium nanoparticles supported onto

APTES functionalized silica colloids after the first cycle of heck reaction

(b)……………………………………………………………………………112

Figure 4-7. TEM image showing the palladium nanoparticles supported onto

APTES functionalized silica colloid suspensions after the second cycle of heck reaction (a) Size distribution of palladium nanoparticles supported onto

APTES functionalized silica colloids after the second cycle of heck reaction

(b)……………………………………………………………………………113

Figure 4-8. TEM image showing the palladium nanoparticles supported onto

APTES functionalized silica colloid suspensions after the third cycle of heck reaction (a) Size distribution of palladium nanoparticles supported onto

APTES functionalized silica colloids after the third cycle of heck reaction

(b)…………………………………………………………………………....113

Figure 4-9. Comparisons of the catalytic activity obtained with functionalized and unfunctionalized CSMNs……………………………………………….114 xiii

Figure 4-10. Comparisons of the catalytic activity with fresh and reused catalysts of palladium nanoparticles supported onto unfunctionalized silica colloids………………………………...…………………………………….114

Figure 4-11. Comparisons of the catalytic activity with fresh and reused catalysts of palladium nanoparticles supported onto functionalized silica colloids………………………………………………………………………115 xiv

Chapter 1 : INTRODUCTION TO NANOCATALYSIS

Catalyst is a substance that increases the rate of a reaction by lowering the activation energy (Ea) that is required to convert reactants into products. By using a catalyst, the desired product can be obtained faster than without the presence of the catalyst. They are used to make chemicals, medicines, fuels and many other important products. Catalysts also play a major role in improving the quality of our life from environmental perspective. Catalysts containing platinum, palladium, rhodium and other metals are used to control the emission levels of toxic gases emitted from automobiles.

Only a portion of catalysts used today can be termed as nanocatalysts. The small particle size leads to many unique properties of the nanoparticles. The term nanocatalyst is defined as a material that has catalytic properties on at least one nanoscale dimension. They possess immense surface-to-volume ratio because of their small size, which is why they act as attractive catalysts. They also have potential applications in the field of photochemistry, nanoelectronics, optics etc. Although the term nano was used until recent times, researchers traditionally focused on producing very small particles of active catalytic agents in order to maximize the reaction efficiency and to reduce the overall cost of the chemical process. Recent advances in synthetic methods for the production of nanomaterials has produced new nanocatalysts with novel properties and reactivity.

Metal nanoparticles are attractive catalysts compared to the bulk materials due to high surface to volume ratio. Because of their small size, the nanoparticles will have

1

more percentage of atoms on the surface, which leads to increased catalytic activity.

However, the problem associated with these catalysts is that they tend to aggregate at higher temperatures and harsh reaction conditions. In most of the cases, aggregation leads to loss of properties associated with the colloidal transition metal nanoparticles.

During catalysis, aggregation or coagulation of metal nanoparticles leads to significant loss of catalytic activity. The stabilization of the metallic nanoparticles in colloidal solution and the means to preserve the catalytic activity is an important aspect to be considered while synthesizing these catalysts.

Synthesis of Colloidal Metal Nanoparticles

There are many ways to synthesize metal nanoparticles and can be mainly divided into two categories: top down approach and bottom up approach. In the top down process, the bulk materials are used as the starting materials and treated by physical means such as mechanical grinding 1 , mechanical alloying 2 and sputtering techniques 3-6 etc. to synthesize nanomaterials. The particles synthesized by this method generally have a size distribution that is broad.

6 The metal nanoparticles synthesized by this method are typically larger and cannot be reproduced resulting in irreproducible catalytic activity. In the bottom up process, the single atoms or ions are allowed to grow into clusters or nanoparticles using wet chemical synthetic methods such as chemical reduction of metal salts 7, 9 and the decomposition of precursors using thermal 10, 11 , photolytic 8 or sonochemical treatment.

12, 13 The particles synthesized by this method have narrow size distribution of the particles.

2

The method used for the synthesis of metal nanoparticles has control over the size and shape of the particles. To have better control over the particle size and shape, chemical synthesis methods are found to be more suitable than physical synthesis methods. Three major routes used to synthesize transition metal nanoparticles include chemical reduction of metal salts, thermal and photochemical decomposition of metal complexes and electrochemical reduction of metal salts.

Chemical Reduction

The chemical reduction of metal salts is one of the most widespread methods used for the synthesis of transition metal nanoparticles.

7, 9 In this method, the metal salts dissociate to form metal ions. The metal ions are then reduced by a reducing agent to form metal nanoparticles that are stabilized by a stabilizer. Most commonly used reducing agents include solvents such as alcohol 14, 15, 17 , salts such as sodium borohydride 18-26 or sodium citrate 27-29 and gases such as carbon monoxide 33 and hydrogen.

30-32

The alcohols mainly used for the reduction of metal salts include ethanol, methanol and isopropanol. The alcohols act both as a reducing agent and solvent. The transition metal nanoparticles are formed by the reduction of metal salts in the refluxing alcohol and the alcohols are oxidized to form corresponding carbonyl compounds (for example, methanol to formaldehyde, ethanol to acetaldehyde). The size distribution of the metal nanoparticles is dependent on the structure and quantity of the alcohol, the stabilizing agent, metallic precursor, temperature and the base used.

14-17

3

Miyake and co-workers reported that the size of palladium nanoparticles can be controlled by varying the amount of the protective polymer, PVP and the kind or concentration of the alcohol used.

14, 15 The smaller sized palladium and platinum nanoparticles were obtained in case of reduction of H

2

PdCl

4

or H

2

PtCl

4

with increase in the amount of the polymer used.

14, 15 Also, the particle size and standard deviation became smaller when the alcohol used has higher boiling point.

15 It was also reported that smaller size of the particles is observed during the reduction of H

2

PtCl

6 or PdCl

2 in the presence of sodium hydroxide and methanol.

17

Borohydrides are the most commonly used reducing agents to synthesize colloidal transition metal nanoparticles. Sodium borohydride is dissolved in aqueous solution and used immediately to avoid decomposition of borohydride to borane and gaseous hydrogen that could escape from the solution. Sodium borohydride is also used as a reducing agent to synthesize platinum and palladium nanoparticles stabilized by PAMAM dendrimers.

22 Phosphine dendrimer stabilized nickel nanoparticles 20 and dendrimer encapsulated iron-oxide nanoparticles 19 were synthesized using sodium borohydride as reducing agent. PEGylated dendrimer stabilized gold nanoparticles were synthesized using sodium borohydride as reducing agent in methanol.

26

Sodium citrate is used as a reducing agent to synthesize Ir and Pt nanoparticles.

28-29 Turkevitch and co-workers used sodium citrate as reducing agent to synthesize Au nanoparticles.

27 The citrate anion acts both as a reducing agent and stabilizer. The decrease in the amount of sodium citrate results in larger size particles.

The reduction in the amount of sodium citrate will result in less number of citrate ions

4

available for stabilizing the particles. This causes the aggregation of smaller particles to form bigger ones.

27

The hydrogen gas is another commonly used reducing agent to synthesize colloidal transition metal nanoparticles.

30-32 This method involves bubbling of hydrogen gas into a solution containing precursor metal salt and a slow reduction process allows the formation of colloidal metal nanoparticles. Hydrogen gas is used as a reducing agent to synthesize Au, Ag, Pt, Pd, Ir, Rh and Ru colloids in the presence of polyvinylalcohol.

30 Polyphosphate is used as a stabilizer to synthesize Ag nanoparticles using hydrogen as reducing agent.

32 Carbon monoxide is used as a reducing agent to synthesize Au nanoparticles in the presence of polyvinylsulfate.

33

Thermal and Photochemical Decomposition

Thermal decomposition involves the decomposition of organometallic salts to respective zero-valent species in high boiling point solvents. Meguro and co-workers reported the synthesis of palladium nanoparticles by heating a solution of palladium acetate to 110 0 C.

34 Using this method, broad size distribution of particles is obtained if secondary stabilizers are not used.

Photochemical decomposition of metal salts was found to be a promising method for the synthesis of monodisperse transition metal nanoparticles. Using this method, metal salts are reduced by radiolytically produced reducing agents or radio induced degradation of organometallic complexes. The main advantage of this method is that a large number of atoms are homogeneously and instantaneously produced during irradiation. This allows the formation of monodisperse nanoparticles. Au, Ag,

5

Pd, Pt, Cu, and Ir nanoparticles have been synthesized by this method.

35-38 During radiolysis of transition metal salts in aqueous solution, solvated electrons or radicals are formed which react with other molecules in solution to form new radicals. These solvated electrons and newly formed radicals act as reductant to reduce the metal salts forming the nanoparticles. The ionization radiation sources used include UV-Vis, x- ray or !

-ray generators. Au, Ag, and Pt nanoparticles have been synthesized by irradiating the corresponding transition metal salt with UV-Visible light in the presence of polymers or surfactants.

39-41

The irradiation time used was quite long using photochemical decomposition methods. To solve this problem, Scaiano and co-workers developed methods in which the irradiation time was reduced to several minutes.

42-45 Ag nanoparticles were synthesized using hydroxydiphenylmethyl as a source of radical species and the irradiation time required to reduce the metal salt to Ag was decreased to few minutes.

42

Electrochemical Reduction

Reetz et al developed electrochemical method for the synthesis of monodisperse transition metal nanoparticles.

46, 47 The mechanism proposed by authors include the following steps: oxidative dissolution of anode to form metal ions, migration of metal ions to cathode, reduction of metal ions at the cathode, nucleation and growth of metal particles, and stabilizing the metal nanoparticles with protecting agents such as tetraalkyl ammonium salts. Adjusting the distance between electrodes, reaction time, temperature and the polarity of solvent can control the size of the nanoparticles. The

6

particle size can also be controlled by the current intensity. The increase in current intensity resulted in smaller transition metal nanoparticles. Pd, Ni, Co, Fe, Ag and Au nanoparticles were synthesized using this method.

46, 47

Stabilization of Metal Nanoparticles

Transition metal nanoparticles in colloidal solution have been used as catalysts for a variety of organic reactions because of their excellent catalytic activity. In general, it is difficult to separate and recycle the catalyst from the suspension of the reaction mixture. Furthermore, they tend to aggregate or coagulate which leads to the loss of catalytic activity. Hence, in order to prevent the aggregation of particles, they have to be stabilized using stabilizers. The selection of the stabilizers is important while synthesizing transition metal nanoparticles. The size and shape of the metal nanoparticles depend on the chemical structure of the stabilizer used. Based on the nature of protective agents used, it is possible to identify five different kinds of stabilization. (i) electrostatic stabilization due to adsorbed ions, (ii) steric stabilization due to the presence of bulky groups, (iii) electrosteric stabilization by the presence of surfactants, (iv) stabilization by ligands and (v) stabilization by dendrimers.

Electrostatic Stabilization

The stabilization of transition metal nanoparticles can be achieved by means of electrostatic stabilization. Ionic compounds such as halides, carboxylates and polyoxoanions when dissolved in solution can produce electrostatic stabilization. The

7

adsorption of these compounds and their related counter ions on the metallic surface will generate an electrical double layer around the particles. This results in Coulombic repulsion between the nanoparticles that prevents the aggregation of the particles.

48-51

Steric Stabilization

The stabilization of metal nanoparticles can also be achieved by the use of macromolecules such as polymers or oligomers.

52, 53 The steric stabilization of metal nanoparticles can be achieved by the steric bulk framework of polymers.

52 Most commonly utilized polymers include PVP 54-56, 58, 59 , polyacrylate 57, 60, 61 , polystyrene 62,

63 etc. Polyacrylate has also been used to stabilize cubic shaped platinum nanoparticles 60 . The bulky groups of PVP, polyacrylate and polystyrene will form a protective layer around the surface of nanoparticles thereby preventing the aggregation of the particles.

Of all the polymers, PVP is mostly used because of its low cost and its ability to stabilize nanoparticles with accessible surface areas in polar solvents such as water.

PVP do not completely passivate the nanoparticle surface. It was speculated that one part of PVP chain adsorbs to the surface of nanoparticle and other part suspends freely in solution forming a protective layer.

58 One of the disadvantage associated with the

PVP is the problematic separation of nanoparticles from the polymer.

64, 65

PVP has been used to stabilize platinum 56, 57 , palladium 54, 55, 64 and rhodium 59 spherical shaped nanoparticles. PVP stabilized nanoparticles have been used as catalyst for various reactions. El-Sayed and co-workers reported that palladium nanoparticles stabilized by PVP acts as efficient catalysts for Suzuki cross-coupling

8

reactions.

67 They reported that by varying the PVP/metal ratio, the size of the nanoparticle can be varied.

66 They have also shown that the decrease in palladium nanoparticle size down to 3 nm improved the catalytic activity in the Suzuki reaction.

64, 67

Electrosteric Stabilization

The electrostatic and steric stabilization can be provided by the use of ionic surfactants such as polyoxoanions.

68, 69 These compounds have a polar head group that generates electrical double layer and lypophilic side chain that provide steric repulsion. Surfactant such as ammonium (Bu

4

N + )/polyoxoanion has bulky ammonium counterions with a highly charged polyoxoanion that provide electrosterical stabilization towards agglomeration.

68, 69

Stabilization by Ligands

The stabilization of metal nanoparticles can also be achieved by the use of ligands such as phosphines 53, 70-71 , thiols 53, 72, 73 or amines 53 etc. The stabilization occurs by coordination of nanoparticles with ligand molecules. Platinum, palladium and gold nanoparticles in colloidal solution have been stabilized using phosphine ligands.

70, 71 Gold nanoparticles have been stabilized using alkanethiols and functionalized thiols that were referred to as monolayer-protected clusters (MPCs).

72,

73 Thiol stabilized nanoparticles are often shown to be less catalytically active or inactive due to passivation of metal surface. Astruc and coworkers synthesized palladium nanoparticles stabilized by dodecanethiolate ligands.

74 Thiol stabilized

9

palladium nanoparticles are reported to be stable and recyclable catalysts for the

Suzuki reaction of aryl halides under ambient conditions.

Stabilization using Dendrimers

Recently, dendrimers are used as stabilizers to synthesize transition metal nanocatalysts.

18, 22-24, 75-76 Dendrimers are a class of polymeric materials that are highly branched monodisperse macromolecules. Dendrimers become densely packed as the chain grows and forms a closed membrane-like structure. Most commonly used dendrimers include PAMAM poly(amido amine) 76 and PPI poly(propylene imine) dendrimers 75 . The nanoparticles synthesized by this method do not agglomerate and are stabilized by encapsulation within the dendrimer.

Crooks and co-workers synthesized Cu, Au, Pt, Pd, Fe, and Ru nanoparticles using poly(amidoamine) (PAMAM) dendrimers.

22, 23 Dendrimers with higher steric bulk play an important role in the stability and catalytic activity exhibited by the dendrimer encapsulated nanoparticles. El-Sayed and co-workers have shown that

Generation 3 and Generation 4 PAMAM dendrimers act as good stabilizers for palladium nanoparticles as higher generation dendrimers encapsulate the nanoparticles more strongly.

77 They also reported that fourth generation PAMAM-OH dendrimer encapsulated palladium nanoparticles being more stable than PVP-Pd nanoparticles and can be recycled several times before aggregating out of solution.

78

10

Supported Nanoparticle Catalysts

The smaller nanoparticles (< 10 nm) are thermodynamically unstable because of their high surface energies and large surface areas. It is difficult to stabilize the nanoparticles by maintaining the small size range while retaining sufficient catalytic activity. Also, the separation of catalyst from unused reactants and products at the end of the reaction is difficult. To overcome this problem, transition metal nanoparticles can be supported onto a heterogeneous solid support.

In heterogeneous catalysis, transition metal nanoparticles supported on various substrates such as silica, carbon and alumina are used as catalysts. Supported metal nanoparticles have been used as catalysts for a variety of industrially important catalytic reactions such as hydrogenations 79, 80 , dehydrogenations 81-83 and oxidation reactions.

80 The supported metal nanocatalysts have advantages such as enhanced activity, selectivity and prevention of aggregation by immobilization on solid support.

Several strategies have been developed for the synthesis of supported metal nanoparticles. Some of the major routes used for the synthesis of supported metal nanocatalysts are described below.

The impregnation method is mostly used method for the synthesis of supported metal catalysts. This process involves wetting of solid support with the precursor metal solution. In the first step, metal salts are dissolved in minimum amount of solvent, followed by the addition of solid support to form slightly wet powder. The solvent is then removed by drying and calcination under oxidizing and reducing

11

conditions.

84 The supported catalysts synthesized by this method are polydisperse and the dispersity depends on the metal, metal loading and support.

85, 86

Deposition/Precipitation or co-precipitation is another way that is used to synthesize supported metal catalysts.

87 This method involves dissolution of precursor metal salt in appropriate solvent, followed by adjustment of pH to achieve complete precipitation of precursor. The precipitate is then deposited on the surface of support followed by calcination. Highly dispersed gold catalysts on different supports have been synthesized using this method.

89 The size distribution of nanoparticles and the degree of dispersion depends on the nature of the support, pH and concentration of the precursor solution, the temperature conditions used for drying and calcination.

87

Grafting is another method that is used to synthesize supported metal catalysts.

90, 91 This method involves formation of covalent bond between the metal precursor and the functional groups on the support. The supports that are used for grafting include polyacrylamide gels 92-95 and polystyrene microspheres 96, 97 etc.

Platinum and rhodium colloids have been grafted onto polyacrylamide gel having aminoethyl groups.

91-93

Microwave irradiation exposure is another method that is used synthesize supported metal catalysts.

98-100 Silver nanoparticles are reduced on graphene oxide under microwave irradiation conditions using starch as a stabilizer.

98 Gold, silver and palladium nanoparticles are supported on mesoporous silica by microwave irradiation of metal salts.

99 These catalysts were found to be highly active in oxidation reactions.

This method allows shorter reaction times compared to the conventional heating

12

methods that take longer. This method also allows the formation of smaller particles and narrow particle size distributions.

100

Supports

The supports used for the synthesis of supported metal nanocatalysts include a variety of solid materials. The solid materials generally used for the synthesis of supported metal catalysts include carbon, silica, alumina, titania 101-109 etc.

Transition metal nanoparticles supported on carbon have been used for a variety of heterogeneous catalytic applications. Carbon materials are most widely used as supports for metal nanoparticles because of their low cost, thermal and chemical stability. The surface of the carbon can be tuned using different approaches such as acid base treatment or heat treatment to optimize catalyst support interactions.

Different kinds of carbon supports used include carbon nanotubes 108, 109 , carbon black 111 , diamond 110 , nanoporous carbon etc. Activated charcoal is most widely used carbon support for the synthesis of supported metal nanocatalysts.

112, 113 Gold, palladium, platinum and bimetallic platinum-ruthenium nanoparticles have also been supported on carbon.

114

Transition metal nanoparticles supported on metal oxides have been used as catalysts for a variety of organic reactions. Metal oxides supports have high surface areas, high thermal stabilities and organized pore structures. Silica 103, 104 , alumina 106,

107 and titania 105, 106 are the most commonly used metal oxide supports. Different kinds of silica supports used include silica gel 125 , silica monoliths 115, 116 , mesoporous

13

silica 103, 104, 135 etc. Silica supported transition metal nanoparticles have been used to catalyze reactions such as hydrogenations 117, 118 , Heck cross coupling reactions 119, 120 etc.

Nanocatalysts for Suzuki and Heck Cross-Coupling Reactions

In recent times, carbon-carbon bond forming reactions discovered by Noble laureates Suzuki, Heck and Negishi have been catalyzed using a wide variety of homogeneous and heterogeneous nanocatalysts.

121 The homogeneous catalysts were used because of their high activity and selectivity, however it is difficult to separate the catalyst at the end of the reaction. It is important to separate the catalyst and reuse it several times to recover the cost of metal in reasonable time. Catalysts that are active but difficult to recover are not generally preferred in chemical industries. To overcome this issue associated with homogeneous catalysts lot of research is done on synthesizing heterogeneous nanocatalysts where the metal is loaded on solid supports such as silica, alumina, carbon etc. However, the active sites on the solid support are not easily accessible to the reactants resulting in decreased catalytic activity. Therefore it is important to synthesize catalysts that are highly active, stable, easily recoverable and reusable.

Suzuki Reaction

The Suzuki cross-coupling reaction is the reaction of aryl or vinyl boronic acids with aryl or vinyl halides in the presence of palladium that result in the

14

formation of biaryls. Reetz et al used Pd, Pd/Ni clusters stabilized by tetraalkyl ammonium salts or PVP as catalysts for Suzuki reaction.

122 They observed that bromo aromatics showed better catalytic activity compared to chloro aromatics. El-Sayed and co-workers catalyzed the Suzuki reaction on aryl iodide and phenylboronic acid using palladium nanoparticles stabilized by PVP in water.

123 They observed the precipitation of palladium black during the reaction.

Rothenburg and co-workers synthesized Pd, Pt, Cu, Ru and mixed bimetallic nanoparticles and used them as catalysts for Suzuki reaction between phenylboronic acid and iodobenzene in DMF using K

2

CO

3 as a base.

124 Palladium nanoparticles showed higher activity among the monometallic catalysts and no reaction was observed with platinum nanoparticles. But, ruthenium and copper showed some catalytic activity and are found to be stable. Among the bimetallic ones, Cu/Pd showed most activity on par with palladium nanoparticles.

124

The problem associated with these homogeneous catalysts is difficulty recycling and metal contamination. The heterogenization of homogeneous catalysts is extensively being done in recent years by incorporation of metal nanoparticles in inorganic supports such as silica, carbon, zeolites etc. Palladium nanoparticles immobilized on PEG functionalized silica gels were used as heterogeneous catalysts for Suzuki reaction and are recycled several times without much loss in catalytic activity.

125 Palladium nanoparticles embedded in carbon thin film lined SBA-15

(mesoporous silica) 126 were used as heterogeneous catalysts for Suzuki reaction between iodobenzene and phenylboronic acid. It was found that the catalyst can be reused five times without any significant loss in activity.

15

Heck Reaction

The palladium catalyzed reaction between an aryl halide and alkene was first reported by Mizoroki 127 and Heck 128 in early 1970s. The Mizoroki-Heck reaction has been most widely used for the carbon-carbon bond formation reactions that allow alkylation, arylation, vinylation of various alkenes through their reaction with alkyl, aryl and vinyl halides in the presence of palladium and base in a single step reaction.

This reaction is mostly carried out in the presence of palladium catalysts involving phosphine ligands. But, these ligands are air-sensitive, poisonous and unrecoverable from the reaction medium. Later, researchers tried to test phosphine-free catalytic systems and were successful.

129, 130

Beller and Reetz for the first time showed that the Heck and Suzuki reaction can also be catalyzed by preformed palladium nanoparticles that are stabilized by tetraalkyl ammonium halides or PVP.

131, 132 But, the problems with these homogeneous catalysts is extreme difficulty of separation, recycling, loss of metal or deactivation via aggregation of palladium nanoparticles formed in situ during heck reaction. To overcome these problems, many chemists have tried designing heterogeneous catalytic systems. This system allows easy catalyst recovery, product separation and inhibits loss of metal. In this regard, the homogeneous palladium nanoparticles are supported or immobilized on PVP-grafted silica 133 , PVI grafted silica 134 and amine functionalized mesoporous silica 135 and are used as catalysts for

Heck reaction. All the studies have shown leaching of palladium metal into the solution from the support.

16

Homogeneous and Heterogeneous Catalysis

In homogeneous catalysis, the catalyst remains in the same phase as the reactants and products. In this type of catalysis, colloidal transition metal nanoparticles finely dispersed in organic or aqueous solutions are used. In heterogeneous nanocatalysis, the catalyst and the reactants are in different phases. Usually, the nanocatalyst will be a solid and the reactants will be either liquids or gases.

Traditionally available heterogeneous nanocatalysts include metal nanoparticles adsorbed onto bulk supports such as silica, alumina or carbon.

101-109 The main advantage associated with homogeneous nanocatalysts is that they have high selectivity compared to the heterogeneous nanocatalysts. The disadvantage associated with homogeneous catalysts includes difficulty recycling the catalyst, metal contamination and poor thermal stability. The heterogeneous catalysts have good thermal stability and the catalyst can be recovered and recycled several times before losing its activity.

The main goal of our research is to synthesize colloidal supported metal nanoparticles (CSMNs) that acts as an intermediate between heterogeneous and homogeneous catalysts. The support that we used is silica colloids and the active catalyst is palladium nanoparticles. The CSMNs have a high surface area and highly active surface atoms. They have the ability to be suspended in solution during catalysis and can be easily separated from the reaction mixture. The CSMNs combine the advantages of heterogeneous catalysts in a near homogeneous format.

17

In Chapter 2 we focused on synthesizing Colloidal Supported Metal

Nanoparticles. The CSMNs consists of palladium nanoparticles covalently attached to the surface of functionalized silica colloids. The surface of the silica colloids is functionalized with 3-mercaptopropyl trimethoxysilane (MPTMS) or 3-aminopropyl triethoxysilane (APTES). The CSMNs are characterized using TEM and EDS. The stability and catalytic activity exhibited by the CSMNs is investigated. For this purpose, the CSMNs are used to catalyze the Suzuki cross-coupling reaction between phenylboronic acid and iodobenzene that forms biphenyl. The amount of biphenyl formed is quantitatively analyzed using HPLC.

In Chapter 3 we focused on investigating the catalytic activity obtained for three types of nanocatalysts: colloidal supported metal nanoparticles (CSMNs) prepared with silica colloids in solution that are air-dried, CSMNs prepared with dry silica colloids resuspended in doubly deionized water that are air-dried, and palladium nanoparticles loaded onto bulk silica dispersed in doubly deionized water that are airdried. The three types of catalysts are prepared with and without the aminopropyltriethoxysilane (APTES) linker. All the six different types of catalysts are characterized using TEM and EDS. The catalytic activity exhibited by six different catalysts is tested on Suzuki cross-coupling reaction between phenylboronic acid and iodobenzene that results in the formation of biphenyl.

In Chapter 4 we focused on investigating the catalytic activity, stability and recycling potential of two different types of nanocatalysts: The palladium metal nanoparticles that are adsorbed onto the surface of unfunctionalized silica colloids, palladium nanoparticles covalently attached to the surface of functionalized silica

18

colloids. We used aminopropyltriethoxysilane (APTES) as the linker to functionalize silica colloids. Both these catalysts are used to catalyze Heck cross-coupling reaction between styrene and iodobenzene to form transstilbene. The fresh catalysts and recycled catalysts are characterized using TEM to observe any changes that might be happening to the catalysts after its use in the reaction.

19

References

1) Gubin, S. P.; Koksharov, Y. A.; Khomutov, G. B.; Yurkov, G. Y. Russian chemical reviews . 2005 , 74 (6) 489-520

2) Gaffet, E.; Tachikart, M.; Kedim, O. E.; Rahouadj, R.; Mater.Charact.

1996 ,

36 , 185-190

3) Chung, B. X.; Liu, C. P.; Materials Letters . 2004 , 58 , 1437-1440

4) Asanithi, P.; Chaiyakun, S.; Limsuwan, P.; Journal of Nanomaterials . 2012 ,

2012 , 963609, 8 pages

5) Suzuki, S.; Suzuki, T.; Tomita, Y.; Hirano, M.; Okazaki, K.; Kuwabata, S.;

Torimoto, T.; Cryst. Eng. Comm . 2012 , 14 , 4922-4926

6) Bouchat, V.; Feron, O.; Gallez, B.; Masereel, B.; Michiels, C.; Borght, T. V.;

Lucas, S.; Surface & Coatings Technology.

2011 , 205 , S577-S581

7) Guzmán, M. G.; Dille, J.; Godet, S.; International Journal of Chemical and

Biological Engineering. 2009, 2:3 , 104-111

8) Sato-Berr´u, R.; Red´on, R.; V´azquez-Olmos, A.; Saniger, J.; J. Raman

Spectrosc. 2009 , 40 , 376-380

9) Khanna, P. K.; Journal of Synthesis and Reactivity in Inorganic, Metal-

Organic, and Nano- Metal Chemistry . 2007 , 37, 805-808

10) Logvinenko, V.; Polunina, O.; Mikhailov, Y.; Mikhailov, K.; and Bokhonov,

B.; Journal of Thermal Analysis and Calorimetry.

2007 , 90, 813-816

11) Li, Y.; Liu, J.; Wang, Y.; Wang, Z. L.; Chem. Mater.

2001 , 13, 1008-1014

12) Bunker, C. E.; Novak, K. C.; Guliants, E. A.; Harruff, B. A.; Meziani, M. J.;

Lin, Y.; Sun,Y. P. Langmuir. 2007 , 23 , 10342-10347

13) Fernando, K. A. S.; Smith, M. J.; Harruff, B. A.; Lewis, W. K.; Guliants, E.

A.; Bunker, C. E. J. Phys. Chem. C. 2009 , 113 , 500-503

14) Teranishi, T.; Miyake, M. Chem. Mater. 1998 , 10 , 594

15) Teranishi, T.; Hosoe, M.; Miyake, M. Adv. Mater. 1997 , 9 , 65

16) Bonet, F.; Delmas, V.; Grugeon, S.; Urbina, H. R.; Silvert, P.; Tekaia-

Elhsissen, K. Nanostruct. Mater . 1999 , 11 , 1277-1284

20

17) Yu, W.; Liu, M.; Liu, H.; Zheng, J. J. Colloid Interface Sci. 1999 , 210 , 218

18) Guoping, L.; Wenting, L.; Yunjun, L.; Min, X.; Chunpeng, C.; Xiaoqing, W.

Chin. J. Chem. 2012 , 30 , 541-546

19) Baykal, A.; Toprak, M. S.; Durmus, Z.; Sozeri, M. S. H.; Demir, A. J.

Supercond. Nov. Magn.

2012 , 25 , 1541-1549

20) Wu, L.; Ling, J.; Wu, Z.; Adv. Synth. Catal. 2011 , 353 , 1452-1456

21) Balogh, L.; Tomalia, D. A. J. Am. Chem. Soc. 1998 , 120 , 7355

22) Zhao, M.; Crooks, R. M. Angew. Chem., Int. Ed. Engl. 1999 , 38 , 364

23) Zhao, M.; Sun, L.; Crooks, R. M. Polym. Prepr . 1999 , 40 , 400

24) Chechik, V.; Crooks, R. M. J. Am. Chem. Soc. 2000 , 122 , 1243

25) Zhao, M.; Crooks, R. M. Adv. Mater. 1999 , 11 , 217

26) Boisselier, E.; Diallo, A. K.; Salmon, L.; Ornelas, C.; Ruiz, J.; Astruc, D.; J.

Am. Chem. Soc.

2010 , 132 , 2729–2742

27) Turkevich, J.; Stevenson, P. C.; Hillier, J. Discuss. Faraday Soc. 1951 , 55-75

28) Harriman, A.; Thomas, J. M.; Millward, G. R. New J. Chem. 1987 , 11 , 757-

762

29) Furlong, D. N.; Launikonis, A.; Sasse, W. H. F.; Sanders, J. V. J. Chem. Soc.,

1984 , 80 , 571-588

30) Tan, C. K.; Newberry, V.; Webb, T. R.; McAuliffe, C. A. J. Chem.Soc., Dalton

Trans. 1987 , 6 , 1299

31) Yu, W.; Liu, H. Chem. Mater. 1998 , 10 , 1205

32) Ershov, B. G.; Abkhalimov, E. V.

Colloid Journal, 2007 , 69

33) Kopple, K.; Meyerstein, D.; Meisel, D. J. Phys. Chem. 1980 , 84 , 870

34) Tano, T.; Esumi, K.; Meguro, K. J. Colloid Interface. Sci. 1989 , 133 , 530-533

35) Ershov, B. G.; Janata, E.; Michaelis, M.; Henglein, A. J. Phys. Chem. 1991 ,

95 , 8996-8999

36) Mosseri, S.; Henglein, A.; Janata, E. J. Phys. Chem. 1989 , 93 , 6791-6795

21

37) Marignier, J. L.; Belloni, J.; Delcourt, M. O.; Chevalier, J. P. Nature 1985 ,

317 , 344-345

38) Michaelis, M.; Henglein, A. J. Phys. Chem. 1992 , 96 , 4719-4724

39) Toshima, N.; Takahashi, T.; Hirai, H. Chem. Lett. 1986 , 35-38

40) Torigoe, K.; Esumi, K. Langmuir. 1992 , 8 , 59-63

41) Yonezawa, Y.; Sato, T.; Ohno, M.; Hada, H. J. Chem. Soc., Faraday Trans.

1987 , 83 , 1559-1567

42) Scaiano, J. C.; Aliaga, C.; Maguire, S.; Wang, D. J. Phys. Chem. B. 2006 , 110 ,

12856- 12859

43) Gonzalez, C. M.; Liu, Y.; Scaiano, J. C. J. Phys. Chem. C 2009 , 113 , 11861-

11867

44) Marin, M. L.; McGilvray, K. L.; Scaiano, J. C. J. Am. Chem. Soc. 2008 , 130 ,

16572-16584

45) McGilvray, K. L.; Decan, M. R.; Wang, D. S.; Scaiano, J. C. J. Am. Chem.

Soc. 2006 , 128 , 15980-15981

46) Reetz, M. T.; Helbig, W. J. Am. Chem. Soc. 1994 , 116 , 7401-7402

47) Reetz, M. T.; Quaiser, S. A. Angew. Chem. Int. Ed. 1995 , 34 , 2240-2241

48) Hunter, R. J. Foundations of Colloid Science, vol. 1, Oxford University Press,

New York, 1986

49) Hirtzel, C. S.; Rajagopalan, R. Colloidal Phenomena: Advanced Topics ,

Noyes Publications, Park Ridge, NJ, 1985

50) Bard, A. J.; Faulkner, L.R. Electrochemical Methods: Fundamentals and

Applications , Wiley, New York, 1980

51) Hunter, R. J. In Foundations of Colloid Science ; Oxford University Press:

New York, 1987; Vol. 1, p 316

52) Napper, D. H. In Polymeric Stabilization of Colloidal Dispersions ; Academic

Press: London, 1983

53) Astruc, D.; Lu, F.; Aranzaes, J. R. Angew. Chem. Int. Ed. 2005 , 44 , 7852-7872

54) Teranishi, T.; Miyake, M. Chem. Mat.

1998 , 10 , 594

22

55) Li, Y.; Hong, X. M.; Collard, D. M.; El-Sayed, M. A. Org. Lett. 2000 , 2(15) ,

2385

56) Narayanan, R.; El-Sayed, M. A. J. Phys. Chem. B , 2003 , 107(45) , 12416

57) Narayanan, R.; El-Sayed, M. A. J. Phys. Chem. B , 2004 , 108(18) , 5726

58) Hirai, H.; Yakurat, N. Polym. Adv. Tech.

2001 , 12(11-12) , 724

59) Borsla, A.; Wilhelm, A. M.; Delmas, H. Catal. Today . 2001 , 66(2-4) , 389

60) Ahmadi, T. S.; Wang, Z. L.; Green, T. C.; Henglein, A.; El-Sayed, M. A.

Science , 1996 , 272 , 1924

61) Li, Y.; Petroski, J.; El-Sayed, M. A. J. Phys. Chem. B , 2000 , 104(47) , 10956

62) Lakina, N. V.; Sul'man, E. M.; Matveeva, V. G.; Mikhailov, I. A, Pharm.

Chem. J.

2000 , 34(3) , 138

63) Sulman, E.; Lakina, N.; Sulman, M.; Ankudinova, T.; Matveeva, V.; Sidorov,

A.; Sidorov, S., Stud. Surf. Sci. Catal.

2000 , 130B , 1787

64) Narayanan, R.; El-Sayed, M. A. J. Am. Chem. Soc. 2003 , 125 , 8340-8347

65) Pachon, L. D.; Rothenberg, G. Appl. Organometal. Chem. 2008 , 22 , 288-299

66) Li, Y.; El-Sayed, M. A. J. Phys. Chem. B 2001 , 105 , 8938-8943

67) Li, Y.; Boone, E.; El-Sayed, M. A. Langmuir. 2002 , 18 , 4921-4925

68) Aiken III, J. D.; Finke, R. G.; J. Mol. Catal. A: Chem. 1999 , 145 , 1-44

69) Lin, Y.; Finke, R. G. J. Am. Chem. Soc. 1994 , 116 , 8335

70) Duteil, A.; Schmid, G.; Meyer-Zaika, W. Chem. Commun. 1995 , 31-32

71) Amiens, C.; Decaro, D.; Chaudret, B.; Bradley, J. S.; Mazel, R.; Roucau, C. J.

Am. Chem. Soc. 1993 , 115 , 11638-11639

72) Dasog, M.; Scott, R. W. J. Langmuir. 2007 , 23 , 3381-3387

73) Hou, W. B.; Dasog, M.; Scott, R. W. J. Langmuir. 2009 , 25 , 12954-12961

74) Lu, F.; Ruiz, J.; Astruc, D. Tetrahedron Lett. 2004 , 45 , 9443-9445

23

75) Crooks, R. M.; Zhao, M. Q.; Sun, L.; Chechik, V.; Yeung, L. K. Acc. Chem.

Res. 2001 , 34 ,181-190

76) Narayanan, R.; El-Sayed, M. A. Top. Catal. 2008 , 47 , 15-21

77) Li, Y.; El-Sayed, M. A. Journal of Physical Chemistry B . 2001 ,105(37), 8938

78) Narayanan, R.; El-Sayed, M. A. Journal of Physical Chemistry B. 2004 ,

108(25), 8572

79) Bynghoon, Y.; Horng-Bin, P.; Wai, C. M. Journal of Physical Chemistry C.

2009, 113(4), 1520-1525

80) Jacinto, M. J.; Santos, O. H. C. F.; Jardim, R. F.; Landers, R.; Rossi, Liane, M.

Applied Catalysis, A: General.

2009 , 360(2), 177-182

81) Dummer, N. F.; Salem, B.; James, H.; Robert, J.; Hutchings, G. J. Catalysis

Today. 2011 , 160(1), 50-54

82) Satoshi, H.; Yindee, S.; Atsushi, O.; Chiemi, T.; Xiuqin, C.; Shaoming, Y.;

Hideki, S.; Masahiko, A.; Nick, S. Journal of Catalysis. 2012 , 289 , 266-271

83) Mizuki, T.; Rajaram, B.; Seitaro, N.; Yasuhiro, I. Applied Catalysis, A:

General.

2006 , 307(1), 78-84

84) Choi, H.; Al-Abed, S. R.; Agarwal, S.; Dionysiou, D. D. Chem. Mater. 2008 ,

20 , 3649-3655

85) White, R. J.; Luque, R.; Budarin, V. L.; Clark, J. H.; Macquarrie, D. J. Chem.

Soc. Rev. 2009 , 38 , 481-494

86) Campelo, J. M.; Lee, A. F.; Luque, R.; Luna, D.; Marinas, J. M.; Romero, A.

A. Chem. Eur. J. 2008 , 14 , 5988-5995

87) Ertl, G.; Knozinger, H.; Weitkamp, J. Handbook of Heterogeneous Catalysis ;

Wiley/VCH: New York/Weinheim, 1997

88) Masatake, H.; Susumu, T.; Tetsuhiko, K.; Hiroyuki, K.; Genet, M. J.; Bernard,

D.; Journal of Catalysis. 1993 , 144(1), 175-192

89) Susumu, T.; Masatake, H.; Tetsuhiko, K.; Atsushi, U.; Yoshiko, N. Stud. Surf.

Sci. Catal. 1991 , 63, 695-704

90) Neelgund, G. M.; Aderemi, O. Applied Catalysis, A: General. 2011 , 399(1-

2), 154-160

24

91) Ungureanu, S.; Deleuze, H.; Babot, O.; Achard, M.; Sanchez, C.; Popa, M. I.;

Backov, R. Applied Catalysis, A: General. 2010, 390(1-2), 51-58

92) Toshima, N.; Ohtaki, M.; Teranishi, T. React. Polym. 1991 , 15 , 135

93) Hirai, H.; Ohtaki, M.; Komiyama, M. Chem. Lett.

1986 , 3 , 269-72

94) Hirai, H.; Ohtaki, M.; Komiyama, M. Chem. Lett. 1987 , 1 , 149-51

95) Suzuki, K.; Yumura, T.; Mizuguchi, M.; Tanaka, Y.; Chen, C.; Akashi, M .

J.

Appl. Polym. Sci . 2000 , 77 , 2678

96) Chen, C.; Chen, M.; Serizawa, T.; Akashi, M. Chem. Communications .

1998 , 7 , 831-832

97) Chen, C.; Serizawa, T.; Akashi, M. Chem. Mater. 1999 , 11 , 1381

98) Hana, Y.; Luoa, Z.; Yuwenb, L.; Tiana, J.; Zhua, X.; Wanga, L.; Applied

Surface

Science 2013 , 266 , 188-193

99) Campelo, J. M.; Conesa, T. D.; Gracia, M. J.; Jurado, M. J.; Luque, R.;

Marinas, J. M.; Romero, A. A. Green Chem. 2008 , 10 , 853-858

100) White, R. J.; Luque, R.; Budarin, V. L.; Clark, J. H.; Macquarrie, D. J. Chem.

Soc. Rev.

2009 , 38 , 481-494

101) Kwong-Yu, C.; Jie, D.; Jiawen, R.; Shaoan, C.; Ying, T. K. J. Mater.

Chem.

2004 , 14 , 505-516

102) Kwong-Yu, C.; Small Fuel Cells (5th Ed.) 2004 , 305-334

103) Jong-San, C.; Jin-Soo, H.; Sang-Eon, P. Res. Chem. Intermed.

2003 , 29 , 921-

938

104) Chen, C. L.; Mou, C. Y. Nanotechnol. Catal.

2004 , 1 , 313-327

105) Mingshu, C.; Wayne, G. D. Chem. Soc. Re V .

2008 , 37 , 1860-1870

106) Ganapati, Y. D. Catal. Sur v . Asia 2005 , 9 , 117–137

107) Sun, K.; Liu, J.; Nag, N. K.; Browning, N. D. Recent Res. De v .

Mater. Sci.

2003 , 4 , 521-540

108) Laura, R.; Christian, P.; Philippe, S.; Montserrat, G.; Emmanuelle, T.

ChemCatChem, 2011 , 3(4), 749-754

25

109) Liu, Z.; Lee, J. Y.; Chen, W.; Han, M.; Gan, L. M. Langmuir. 2004 , 20(1),

181

110) Nakagawa, K.; Yamagishi, M.; Nishimoto, H.; Ikenaga, N.; Suzuki, T.;

Kobayashi, T.; Nishitani-Gamo, M.; Ando, T. Chem. Mat . 2003 , 15(24), 4571

111) Takasu, Y.; Itaya, H.; Kawaguchi, T.; Sugimoto, W.; Murakami, Y. Stud.

Surf. Sci. Catal.

2003 , 145 , 279

112) Silvio, C.; Paul, M.; Peter, J.; Ken, G.; Christopher, K. J.; Graham, H. J.

Phys. Chem. Chem. Phys. 2003 , 5(6), 1329

113) Albers, P. W.; Konrad, M.; Frost, C. D.; Parker, S. F. Journal of Physical

Chemistry C . 2011 , 115(50), 24485-24493

114) Liu, Z.; Ling, X. Y.; Lee, J. Y.; Su, X.; Gan, L. M. J. Mat. Chem.

2003 ,

13(12), 3049

115) Haifeng, Y.; Qingyi, L.; Feng, G.; Qihui, S.; Yan, Y.; Fuqiang, Z.; Songhai,

X.; Bo, T.; Dongyuan, Z. Advanced Functional Materials. 2005 , 15(8), 1377-

1384

116) Oemer, D.; Olga, S.; Neil, C.; Ozin, G. A.; Journal of Materials Chemistry .

2003, 13(2), 328-334

117) Doris, R.; Claudio, M.; Fierro, J. L. G.; Patricio, R. Journal of the Chilean

Chemical Society.

2012 , 57(4), 1394-1399

118) Oki, M.; Rebrov, E. V.; Angel, B.; de Croon, M. H. J. M.; Schouten, J. C.

Applied Catalysis, A: General. 2009 , 368(1-2), 87-96

119) Mehran, G.; Donya, V.; Zahra, S. Chinese Chemical

Letters. 2012 , 23(8) , 887-890

120) Gianmario, M.; Luca, B.; Claudio, G.; Raimondo, M.; Giovanni, S.;

Salvatore, C. Catalysis Letters. 2009 , 132(1-2), 50-57

121) N. Miyaura, K. Yamada and A. Suzuki. Tetrahedron Lett.

1979 , 20 , 3437

122) Reetz, M. T.; Breinbauer, R.; Wannlnger, K.

Tetrahedron Letters, 1996,

37(26), 4499-4502

123) Li, Y.; Hong, X. M.; Collard, D. M.; El-Sayed, M. A. Org. Lett.

2000 , 2(15),

2385

26

124) Thathagar, M. B.; Beckers, J.; Rothenberg, G. Journal of the American

Chemical. Society . 2002 , 124(40), 11858-11859

125) Dutta, P.; Sarkara, A. Adv. Synth. Catal . 2011 , 353 , 2814 – 2822

126) Zhi, J.; Song, D.; Li, Z.; Lei, X.; Hu, A. Chem. Commun.

2011 , 47 , 10707–

10709

127) Mizoroki, T.; Mori, K.; Ozaki, A. Bull. Chem. Soc. Jpn . 1971 , 44 , 581

128) Heck, R. F.; Nolley, J. P . J. Org. Chem . 1972 , 37 , 2320

129) Alo, B. I.; Kandil, A.; Patil, P. A.; Sharp, M. J.; Siddiqui, M. A.; Snieckus,

V. J. Org.Chem.

1991 , 56 , 3763

130) Wallow, T. I.; Novak, B. M. J. Org. Chem.

1994 , 59 , 5034

131) Beller, M.; Fischer. H.; Kuhlein, K.; Reisinger. C. P.; Herrmann, W. A.;

J.Organomet. Chem . 1996 , 520 , 257

132) Reetz, M.T.; Breinbauer, R.; Wanninger, K.; Tetrahedron Lett , 1996 , 37 ,

4499

133) Tamami, B.; Allahyari, H.; Ghasemi, S.; Farjadian, F.; J.

Organometallic.Chem

. 2011 , 696 , 594-599

134) Tamami, B.; Allahyari, H.; Ghasemi, S.; Farjadian, F.; Sohelia, G.; Iranian

Polymer Journal, 2011, 20, 699-712

135) Peiyu, W.; Qingshan, L.; Jiangong, L. Materials Research Bulletin. 2010 , 45 ,

129-134

27

MANUSCRIPIT- I

Published in Journal of Physical Chemistry C, 2010 , 114 , 6256-6362

Synthesis and Characterization of Colloidal Supported Metal Nanoparticlees (CSMNs) as Potential Intermediate Nanocatalysts

Kalyani Gude and Radha Narayanan *

Department of Chemistry, University of Rhode Island, Kingston, RI, USA, 02881

28

Chapter 2

Synthesis and Characterization of Colloidal Supported Metal Nanoparticles (CSMNs) as Potential Intermediate Nanocatalysts

Abstract

We report the synthesis of a new type of intermediate nanocatalyst material that we term as colloidal supported metal nanoparticles (CSMNs). The CSMNs that we have synthesized consist of palladium nanoparticles covalently attached to functionalized silica colloids. The synthesis process involves four steps, which include the synthesis of the silica colloids, synthesis of the palladium nanoparticles, functionalization of the silica colloids with 3-mercaptopropyl trimethoxysilane

(MPTMS) or 3-aminopropyl triethoxysilane (APTES), and covalent attachment of the palladium nanoparticles onto the functionalized silica colloids. We have characterized the size of the silica colloids and palladium nanoparticles by using transmission electron microscopy (TEM). In addition, we have characterized the attachment of the palladium nanoparticles onto the two types of functionalized silica colloids by using

TEM as well as energy dispersive spectroscopy (EDS). In the case of the CSMNs prepared by using the silica colloids functionalized with MPTMS, we observed that there is significant amount of additional deposits between each of the functionalized silica colloids. This type of additional deposits is significantly diminished in the case of the CSMNs prepared by using silica colloids functionalized with APTES. We have conducted some initial studies to determine at what stage the additional deposits occur and discuss potential sources that give rise to this phenomenon. We have also

29

conducted initial kinetics study and assessment of the stability of the CSMNs after the catalytic process.

Keywords : transition metal nanoparticles, nanocatalysts, silica colloids, palladium nanoparticles, colloidal supported metal nanoparticles (CSMNs)

Introduction

Metal nanoparticles are very attractive catalysts compared to bulk catalytic materials due to their high surface-to-volume ratio. Some types of traditional nanocatalysts include transition metal nanoparticles in colloidal suspension 1-8 , those adsorbed onto bulk supports 9-14 , and lithographically fabricated arrays of nanocatalysts 15-19 . Transition metal nanoparticles in colloidal suspension have been synthesized by using a wide variety of reducing agents 20 such as hydrogen 21-23 , sodium borohydride 24 , and ethanol 25-27 . Many different types of stabilizers have been used as capping agents to stabilize the nanoparticles such as surfactants 28,29 , polymers 29,30 , dendrimers 28,30 , as well as different types of ligands 28,29 . In the case of metal nanoparticles adsorbed onto bulk supports, a wide variety of support materials have been used such as carbon 31,32 , mesoporous silica 33,34 , titania 35,36 , alumina 36,37 , zeolites 9,36 , and resins 36,38 . Arrays of metal nanocatalysts have been fabricated by using electron beam lithography 17,19 as well as colloidal lithographic techniques 39 .

In this paper, we discuss the design of a new type of intermediate nanocatalyst that we term as colloidal supported metal nanoparticles (CSMNs). Some potential advantages of CSMNs as intermediate nanocatalysts include: 1.) being suspended in solution during liquid-phase catalytic reactions, 2.) having high metal loadings on the silica colloid surface resulting in a high surface area, and 3.) facile separation of the

30

reaction mixture from the CSMNs. The CSMNs combine the advantages of heterogeneous catalysts in a near homogeneous format. These advantages make using

CSMNs particularly attractive nanocatalysts for liquid-phase reactions compared to their colloidal counterparts and those adsorbed onto bulk supports. We have utilized a four step process for synthesizing the palladium nanoparticles attached to silica colloids. The four steps in this process include synthesizing the silica colloids, synthesizing the PVP capped palladium nanoparticles, functionalizing the silica colloids, and attaching the PVP capped palladium nanoparticles to the silica colloids.

We have characterized the CSMNs by using TEM and EDS. In the case of the

CSMNs prepared with silica colloids functionalized with 3-mercaptopropyl trimethoxysilane (MPTMS), there is significant amount of additional deposits present between each of the silica colloids. This type of additional deposits is significantly diminished in the case of the CSMNs prepared with silica colloids functionalized with

APTES. We have conducted some initial studies to determine at which stage the additional deposits are formed and discuss potential sources that give rise to this phenomenon.

Experimental

Synthesis of Silica Colloids

The silica colloids were synthesized by using the Stoeber synthesis method 40 .

First, 30 mL of ethanol and 2.4 mL of ammonium hydroxide was added to an

Erlenmeyer flask and stirred for 5 minutes. Then, 1.2 mL of tetraethylorthosilicate

(TEOS) was added to the solution containing ethanol and ammonium hydroxide and

31

this solution was stirred overnight. The solution was initially clear and after 15-20 minutes, the solution starts to turn cloudy and the final solution is very turbid and consists of a suspension of the silica colloids.

Synthesis of PVP Stabilized Palladium Nanoparticles

The palladium nanoparticles were synthesized by using the ethanol reduction method similar to that described previously 25,41-43 . The palladium precursor solution

(H

2

PdCl

4

) was prepared by adding 0.0887 g of PdCl

2

and 6 mL of 0.2 M HCl, and diluting to 250 mL with doubly distilled water. A solution containing 15 mL of 2 mM of H

2

PdCl

4

, 21 mL of doubly deionized water, 0.0667 g of PVP, and 4 drops of 1 M

HCl was heated. When the solution began to reflux, 14 mL of ethanol was added.

The solution was then refluxed for 3 h, and this resulted in a dark brown suspension of

Pd nanoparticles.

Functionalization of Silica Colloids

We have used two different linkers to functionalize the silica colloids: 3mercaptopropyl trimethoxysilane (MPTMS) and 3-aminopropyl triethoxysilane

(APTES). Two separate sets of functionalized silica colloids were prepared in which

100 µ L of MPTMS or APTES is added to the silica colloid suspension.

Synthesis of Colloidal Supported Palladium Nanoparticles

The functionalized silica colloids were first centrifuged four times at 13,500 rpm for 3 minutes each time. During the first two-centrifugation cycles, the

32

functionalized silica colloids were redispersed in ethanol and during the last two centrifugation cycles, they were redispersed in doubly deionized water. Two mL of the centrifuged functionalized silica colloids and four mL of the PVP capped palladium nanoparticles are placed into a scintillation vial and this suspension is mixed for 24 hours to allow the palladium nanoparticles to bind to the functionalized silica colloids.

Characterization Studies with TEM and EDS

One drop of dilute suspensions of the palladium nanoparticles, silica colloids, and the colloidal supported palladium nanoparticles was placed on Formvar coated copper grids. The drop was allowed to air-dry for ~1 hour. The JEOL 2100EX TEM was used to obtain the TEM images of the palladium nanoparticles, silica colloids, and

CSMNs. Energy dispersive spectroscopy (EDS) was used to determine what elements are present in the CSMNs.

Size Distribution Analysis

The UTHSCSA ImageTool for Windows—Version 3 image analysis software was used to determine the size distributions of the silica colloids and the palladium nanoparticles. The Distance tool in the Analysis pull-down menu is used to measure the number of pixels in the scale bar of the TEM image. Based on the number of pixels for the fixed size associated with the scale bar and measuring the number of pixels for ~200 nanoparticles in several TEM images, the size of the nanoparticles can be calculated by dividing the number of pixels for the nanoparticles by the number of

33

pixels of the scale bar and multiplying by the fixed size associated with the scale bar of the TEM image. We then plotted the histogram of % nanoparticles/colloids vs. nanoparticle/colloid size and obtained a Gaussian fit to the histogram. From the

Gaussian fit, we can determine the average size and standard deviation of the palladium nanoparticles and the silica colloids.

Suzuki Cross-Coupling Reaction

The Suzuki reaction between phenylboronic acid and iodobenzene was catalyzed using the PVP-Pd nanoparticles as described previously 25,26,44 . For this reaction, 0.49 g (6 mmol) of sodium acetate, 0.37 g (3 mmol) of phenylboronic acid, and 0.20 g (1 mmol) of iodobenzene was added to 150 mL of 3:1 acetonitrile:water solvent. The solution was heated to 100 o C and 5 mL of the PVP-Pd nanoparticles was added to start the reaction. The reaction mixture was refluxed for a total of 12 hours.

Results and Discussion

Synthesis of Colloidal Supported Metal Nanoparticles (CSMNs)

We discuss the design of a new type of intermediate nanocatalyst that we term as colloidal supported metal nanoparticles (CSMNs). Some important advantages of

CSMNs include: 1.) being suspended in solution during liquid-phase catalytic reactions, 2.) having high metal loadings on the silica colloid surface resulting in a high surface area, and 3.) facile separation of the reaction mixture from the catalyst.

The CSMNs combine the advantages of heterogeneous catalysts in a near-

34

homogeneous format. These advantages make using CSMNs particularly attractive nanocatalysts compared to their colloidal counterparts and those adsorbed onto bulk supports for liquid-phase reactions. Scheme 1 illustrates the four step process that we have used to synthesize the palladium nanoparticles supported onto the silica colloids:

1.) synthesizing the silica colloids, 2.) synthesizing the palladium nanoparticles, 3.) functionalizing the silica colloid surface, and 4.) attaching the palladium nanoparticles onto the silica colloid surface. The silica colloids were synthesized by using the

Stoeber synthesis method and the palladium nanoparticles were synthesized by using the ethanol reduction method and polyvinylpyrolidone as the stabilizer. The silica colloids were functionalized by using two different types of linkers: 3-mercaptopropyl trimethoxysilane (MPTMS) and 3-aminopropyl triethoxysilane (APTES). The palladium nanoparticles were covalently attached to the functionalized silica colloids via the Pd-S bond in the case of the silica colloids functionalized with MPTMS and

Pd-N bond in the case of the silica colloids functionalized with APTES.

The silica colloids were synthesized by using the Stoeber synthesis method 40 and TEM was used to characterize the size of the silica colloids. Based on the size measurements obtained with ImageTool, we have plotted the size distribution histogram and calculated the average size of the silica colloids. Figure 1 shows a representative TEM image of the silica colloids as well as the size distribution histogram obtained based on the size measurements with the ImageTool image analysis software. The size distribution histogram is plotted as % silica colloids vs. silica colloid size. The average size of the silica colloids is 344 + 19 nm and it can be seen that the silica colloids are relatively monodisperse.

35

The palladium nanoparticles were synthesized by using the ethanol reduction method as described previously 25 . Figure 2 shows a representative TEM image of the palladium nanoparticles as well as the size distribution histogram obtained for the palladium nanoparticles. In this case, the size distribution histogram was also plotted as % palladium nanoparticles vs. palladium nanoparticle size. The average size of the palladium nanoparticles is 2.9 + 1.4 nm based on the Gaussian fit of the size distribution histogram.

Figure 3 shows examples of TEM images of the palladium nanoparticles attached to the silica colloids functionalized with MPTMS. We also obtained the energy dispersive spectrum (EDS) of the CSMNs to determine if both Si and Pd peaks were present. The EDS spectrum can give valuable information on the elements present which can help determine whether we have formed CSMNs. As can be seen in Figure 3, there are peaks associated with both Pd and Si that are present which is evidence for the formation of the CSMNs since this is direct evidence of the attachment of the Pd nanoparticles onto the functionalized silica colloids. The silica colloids functionalized with MPTMS have mercapto groups available for the palladium nanoparticles to bind. This results in the Pd-S bond being formed. Also, from Figure 3, it can be seen that there is a significant amount of additional deposits that are present between each of the CSMNs with silica colloids functionalized with

MPTMS. We will discuss this finding in more detail later on in another section of this paper.

We have also investigated the attachment of the palladium nanoparticles by using silica colloids functionalized with APTES. We chose this linker since this

36

would result in the silica colloids being functionalized with amine groups that would be available to bind to the palladium nanoparticles. Figure 4 shows examples of TEM images of CSMNs prepared by attaching the palladium nanoparticles to the silica colloids functionalized with APTES. From the TEM images, it can be seen that the palladium nanoparticles readily attach to the Pd nanoparticle surface since in this case the functionalized silica colloids have amine groups available for the Pd NPs to bind.

The covalent attachment of the palladium nanoparticles occurs by formation of the Pd-

N bond. Also, it can be seen that there is greatly diminished amount of additional deposits observed for the CSMNs prepared with silica colloids functionalized with

APTES. This will be discussed in more detail in the next section of this paper.

Investigations on Additional Deposits Present in CSMNs Prepared with MPTMS

Functionalized Silica Colloids.

As we briefly discussed earlier in this paper, we have observed significant additional deposits in the case of the CSMNs prepared with MPTMS functionalized silica colloids and very little or minimal amounts of these types of deposits in the case of the CSMNs prepared with the APTES functionalized silica colloids. As a result, we have conducted a set of investigations to determine at what stage the deposits are formed. These studies will help determine whether the deposits occurred after the functionalization process or after the stage in which the palladium nanoparticles were allowed to react with the functionalized silica colloids. To answer this question, we have conducted a set of experiments in which we obtained TEM images of the silica colloids before and after functionalization with MPTMS and APTES.

37

Figure 5 and 6 show the TEM images of the silica colloids before and after functionalization with MPTMS and APTES. It can be seen that the deposits are formed on the silica colloid surface after functionalizing the silica colloids with

MPTMS, but not when the silica colloids are functionalized with APTES. As a result, it can be seen that the formation of the deposits occurs after the functionalization process in the case of the MPTMS functionalized silica colloids. This shows that in the case of the CSMNs prepared with silica colloids functionalized with MPTMS, the deposits are formed during the stage in the synthesis where the silica colloids are functionalized with the MPTMS. Also, it can be seen that there are very little or minimal amount of deposits observed after functionalizing the silica colloids with

APTES. This is consistent with the observation that very little or minimal deposits are present in the case of the CSMNs prepared with the silica colloids functionalized with

APTES.

This now raises the question of how these deposits are formed and also why there is significant amount of deposits formed when MPTMS is used as the linker and very little or minimal amount of deposits formed when the APTES is used as the linker. It is worth noting that that the synthesis of the silica colloids occurs by basecatalyzed hydrolysis of tetraethyl orthosilicate (TEOS) 45 . In the Stoeber synthesis method 40 to form silica colloids, ammonium hydroxide is used as the base. It is also worth noting that mesoporous silica materials containing mercaptopropyl or phenyl groups on their surface can be prepared by the sol-gel technique 46 . This technique involves co-hydrolysis of TEOS with MPTMS or phenyltriethoxysilane (PTES) in the

38

presence of hexadecyltrimethylammonium bromide as a templating agent in aqueous sodium hydroxide 46 .

This suggests that in our observations of additional deposits being present in

MPTMS-functionalized silica colloids, there could be a side reaction that also takes place. After synthesizing the silica colloids, 100 microliters of the MPTMS is added directly to the silica colloid suspension. As a result, there could still be some TEOS and ammonium hydroxide also present in the silica colloid suspension. As a result, there are two reactions that can occur when the MPTMS is added to the silica colloid suspension. The main and desired reaction involves the functionalization of the

MPTMS on the surface of the silica colloids and the mercapto groups being available for the palladium nanoparticles to bind to the MPTMS functionalized silica colloids.

The secondary side reaction involves co-hydrolysis of MPTMS with excess unreacted

TEOS still remaining in solution in the presence of the ammonium hydroxide base.

This secondary reaction would result in the formation of amorphous silica materials and would explain the additional deposits that we observed in the TEM images of both the MPTMS functionalized silica colloids and the CSMNs prepared by using MPTMS functionalized silica colloids.

It is also interesting to observe that there are very little or minimal amounts of additional deposits present when the silica colloids are functionalized with APTES and when the palladium nanoparticles are attached to the APTES functionalized silica colloids. It is quite possible that the co-hydrolysis side reaction observed for the

MPTMS functionalized silica colloids occurs at a much lower rate for the APTES functionalized silica colloids. It is worth noting that the co-hydrolysis reaction has

39

been reported to occur with many different silanes including MPTMS, APTES, and

PTES 46,47 . One reason that there is significantly diminished amount of additional deposits present in the APTES functionalized silica colloids and the CSMNs prepared with APTES functionalized silica colloids could be that the rate of functionalization onto the silica colloids is much faster than the rate of the co-hydrolysis process between TEOS and APTES in the presence of ammonium hydroxide base. In the case of MPTMS functionalized silica colloids, the functionalization onto the silica colloids occurs at a much slower rate compared to the co-hydrolysis process between TEOS and MPTMS in the presence of ammonium hydroxide base. Overall, the secondary co-hydrolysis side reaction occurs at a faster rate for the MPTMS functionalized silica colloids compared to the APTES functionalized silica colloids. This would also explain why there is very little or minimal amount of deposits present in the case of the APTMS functionalized silica colloids and a significantly large amount of deposits in the case of the MPTMS functionalized silica colloids. Overall, both types of

CSMNS have the potential to be used as intermediate types of nanocatalysts for liquidphase reactions that require palladium based catalysts.

Catalytic Activity and Stability of the CSMNs

We have used the CSMNs prepared with the amine functionalized silica colloids as catalysts for the Suzuki reaction between phenylboronic acid and iodobenzene to form biphenyl. Reversed phase HPLC was used to follow the kinetics of the biphenyl peak formed during the first hour of the Suzuki reaction between phenylboronic acid and iodobenzene. Figure 7 shows the reaction kinetics during the

40

first hour of the Suzuki reaction. It can be seen that biphenyl starts to form during the first five minutes of the reaction.

We have also examined the stability of the CSMNs after the full course of the

Suzuki reaction (12 hours). Figure 8 shows TEM images of the CSMNs prepared with the amine functionalized silica colloids after the Suzuki reaction as well as the EDS spectra. It can be seen that there are structural changes in the silica colloids that takes place in which the silica colloids become porous. The observed changes in the silica colloids could be due to the high temperatures of the Suzuki reaction and the long reflux times. The palladium nanoparticles continue to be attached to the silica colloid surface as can be seen in the EDS spectrum in which the Pd peaks are present in addition to the Si peak.

Conclusions

We have designed a new type of intermediate nanocatalyst that we have termed colloidal supported metal nanoparticles (CSMNs) that we have characterized by using TEM and EDS. We have prepared the CSMNs by covalently attaching palladium nanoparticles to the MPTMS or APTES functionalized silica colloids. The

CSMNs would serve as an attractive intermediate nanocatalyst in liquid-phase reactions. In the case of the CSMNs prepared with the MPTMS functionalized silica colloids, it is observed that there are additional deposits present between the individual

MPTMS functionalized silica colloids. These additional deposits are formed due to a side-reaction which involves co-hydrolysis of TEOS with MPTMS in the presence of ammonium hydroxide base. The additional deposits are greatly diminished in the case of the APTES functionalized silica colloids and the CSMNs prepared with the APTES

41

functionalized silica colloids and this could be due to the co-hydrolysis side-reaction occurring at a much slower rate. After the full course of the Suzuki reaction, it is observed that there are structural changes in the silica colloids in which it becomes porous and the palladium nanoparticles are still bound to the silica colloids. Overall, both types of CSMNs have the potential to be used as intermediate types of nanocatalysts for liquid-phase reactions that require palladium based catalysts.

Acknowledgments

We thank the University of Rhode Island for the start-up funds that were used to conduct this research. We also thank Dr. Richard Kingsley in the University of

Rhode Island Electron Microscopy Center for assistance in acquiring the TEM images and EDS spectra obtained using the JEOL 2100EX transmission electron microscope.

42

Step 1

Step 2

O

O

Si

O

O tetraethyl orthosilicate

(TEOS)

+ NH

4

OH ammonium hydroxide ethanol silica colloid palladium nanoparticles palladium (II) chloride

+ polyvinylpyrolidone

Step 3

silica colloid

3-mercaptopropyl trimethoxysilane thiol functionalized silica colloid

Step 4

nanoparticles (CSMNs).

+ palladium nanoparticles thiol functionalized silica colloid palladium

CSMNs

Scheme 1.

Four step process for the synthesis of colloidal supported metal

43

b

Average Size of Silica Colloids

= 344 + 19 nm

0 50 100 150 200 250 300 350 400

Size of Silica Colloids (nm)

Figure 2-1.

TEM image of the silica colloids synthesized with the Stoeber method (a) and the size distribution histogram (b) b

20

Average Size

= 2.9 + 1.4 nm

15

10

5

0

0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0

Size of Palladium Nanopartilces (nm)

Figure 2-2.

TEM image of the palladium nanoparticles (a) and the size distribution histogram (b)

10

8

6

4

2

0

44

Figure 2-3.

TEM images and EDS spectrum of CSMNs prepared with silica colloids functionalized with 3-mercaptopropyl trimethoxysilane and palladium nanoparticles

45

Figure 2-4.

TEM image of CSMNs prepared using APTES functionalized silica colloids (a) TEM image of the CSMNs at a higher magnification (b)

Figure 2-5.

TEM images of the silica colloids before and after functionalization with

3-mercaptopropyl trimethoxysilane

46

Figure 2-6.

TEM images of the silica colloids before and after functionalization with

3-aminopropyl triethoxysilane

Pd NPS on SC with 100 µ L Amine Linker

1.0

0.8

0.6

0.4

0.2

1.6

1.4

1.2

0.0

0 10 20 30 40 50 60

Time (min)

Figure 2-7.

Kinetics of the CSMNs prepared with palladium nanoparticles (NPs) bound to silica colloids (SC) functionalized with the amine linker

47

Figure 2-8.

TEM images (a-b) and EDS spectrum (c) of the CSMNs after the Suzuki reaction.

48

References

1) Narayanan, R.; El-Sayed, M. A. J. Phys. Chem. B 2005 , 109 , 12663-12676.

2) Narayanan, R.; El-Sayed, M. A. Chim. Oggi 2007 , 25 , 84-86.

3) Narayanan, R.; El-Sayed, M. A. Top. Catal.

2008 , 47 , 15-21.

4) Narayanan, R.; Tabor, C.; El-Sayed, M. A. Top. Catal.

2008 , 48 , 60-74.

5) Roucoux, A. Top. Organomet. Chem.

2005 , 16 , 261-279.

6) Roucoux, A.; Nowicki, A.; Philippot, K. Nanopart. Catal.

2008 , 349-388.

7) Roucoux, A.; Philippot, K. Handbook of Homogeneous Hydrogenation 2007 ,

1 , 217-256.

8) Wang, D.; Xie, T.; Li, Y. Nano Res.

2009 , 2 , 30-46.

9) Thibault-Starzyk, F.; Daturi, M.; Bazin, P.; Marie, O. Nanopart. Catal.

2008 ,

505-528.

10) Chan, K.-Y.; Chi, N.; Phillips, D. L. Curr. Top. Electrochem.

1998 , 6 , 189-

197.

11) Hermans, S.; Khimyak, T.; Raja, R.; Sankar, G.; Johnson, B. F. G.; Thomas, J.

M. Stud. Surf. Sci. Catal.

2003 , 145 , 181-184.

12) Kumar, R.; Ghosh, A.; Patra, C. R.; Mukherjee, P.; Sastry, M. Nanotech.

Catal.

2004 , 1 , 111-136.

13) Thomas, J. M.; Johnson, B. F. G.; Raja, R.; Sankar, G.; Midgley, P. A. Acc.

Chem. Res.

2003 , 36 , 20-30.

49

14) Baeumer, M.; Libuda, J.; Neyman, K. M.; Roesch, N.; Rupprechter, G.;

Freund, H.-J. Phys. Chem. Chem. Phys.

2007 , 9 , 3541-3558.

15) Somorjai, G. A. Appl. Surf. Sci.

1997 , 121/122 , 1-19.

16) Somorjai, G. A. ACS Symp. Ser.

2005 , 890 , 210-219.

17) Somorjai, G. A.; Rioux, R. M. Catal. Today 2005 , 100 , 201-215.

18) Somorjai, G. A.; York, R. L.; Butcher, D.; Park, J. Y. Phys. Chem. Chem.

Phys.

2007 , 9 , 3500-3513.

19) Eppler, A.; Rupprechter, G.; Guczi, L.; Somorjai, G. A. J. Phys. Chem. B

1997 , 101 , 9973-9977.

20) Burda, C.; Chen, X.; Narayanan, R.; El-Sayed, M. A. Chem. Rev. (Washington,

DC, United States) 2005 , 105 , 1025-1102.

21) Narayanan, R.; El-Sayed, M. A. Nano Lett.

2004 , 4 , 1343-1348.

22) Narayanan, R.; El-Sayed, M. A. J. Amer. Chem. Soc.

2004 , 126 , 7194-7195.

23) Narayanan, R.; El-Sayed, M. A. J. Phys. Chem. B 2004 , 108 , 5726-5733.

24) Antolini, E.; Salgado, J. R. C.; da Silva, R. M.; Gonzalez, E. R. Mat. Chem.

Phys.

2007 , 101 , 395-403.

25) Narayanan, R.; El-Sayed, M. A. J. Amer. Chem. Soc.

2003 , 125 , 8340-8347.

26) Narayanan, R.; El-Sayed, M. A. J. Phys. Chem. B 2004 , 108 , 8572-8580.

27) Narayanan, R.; El-Sayed, M. A. J. Catal.

2005 , 234 , 348-355.

28) Astruc, D.; Lu, F.; Aranzaes, J. R. Ange. Chem., Int. Ed.

2005 , 44 , 7852-7872.

50

29) Wang, Q.; Ostafin, A. E. Encyclopedia of Nanoscience and Nanotechnology

2004 , 5 , 475-503.

30) Mueller, C.; Nijkamp, M. G.; Vogt, D. Eur. J. Inorg. Chem.

2005 , 4011-4021.

31) Chan, K.-Y.; Ding, J.; Ren, J.; Cheng, S.; Tsang, K. Y. J. Mat. Chem.

2004 ,

14 , 505-516.

32) Chan, K.-Y. Small Fuel Cells (5th Edition) 2004 , 305-334.

33) Chang, J.-S.; Hwang, J.-S.; Park, S.-E. Res. Chem. Interm.

2003 , 29 , 921-938.

34) Chen, C. L.; Mou, C. Y. Nanotech. Catal.

2004 , 1 , 313-327.

35) Chen, M.; Goodman, D. W. Chem. Soc. Rev.

2008 , 37 , 1860-1870.

36) Yadav, G. D. Catal. Surv. Asia 2005 , 9 , 117-137.

37) Sun, K.; Liu, J.; Nag, N. K.; Browning, N. D. Rec. Res. Dev. Mat. Sci.

2003 , 4 ,

521-540.

38) Corain, B.; Centomo, P.; Lora, S.; Kralik, M. J. Molec. Catal. A: Chem.

2003 ,

204-205 , 755-762.

39) Wood, M. A. J. Royal Soc., Interf. / Royal Soc.

2007 , 4 , 1-17.

40) Flaschsbart, H.; Stoeber, W. J. Coll. Interf. Sci.

1969 , 30 , 568-573.

41) Li, Y.; Hong, X. M.; Collard, D. M.; El-Sayed, M. A. Org. Lett.

2000 , 2 , 2385.

42) Li, Y.; Boone, E.; El-Sayed, M. A. Langmuir 2002 , 18 , 4921.

43) Teranishi, T.; Miyake, M. Chem. Mat.

1998 , 10 , 594.

51

44) Narayanan, R.; El-Sayed, M. A. Langmuir 2005 , 21 , 2027-2033.

45) Badley, R. D.; Ford, W. T.; McEnroe, F. J.; Assink, R. A. Langmuir 1990 , 6 ,

792.

46) Rac, B.; Molnar, A.; Forgo, P.; Mohai, M.; Bertoti, I. J. Molec. Catal. A:

Chem.

2006 , 244 , 46-57.

47) Rahman, I. A.; Jafarzadeh, M.; Sipaut, C. S. Cer. Int.

2008 , 35 , 1883.

52

MANUSCRIPIT- II

Published in Journal of Physical Chemistry C, 2011 , 115 , 12716-12725

Colloidal Supported Metal Nanoparticles (CSMNs) as Effective Nanocatalysts for

Liquid-Phase Suzuki Cross-Coupling Reactions

Kalyani Gude and Radha Narayanan *

Department of Chemistry, University of Rhode Island, Kingston, RI, USA, 02881

53

Chapter 3

Colloidal Supported Metal Nanoparticles (CSMNs) as Effective Nanocatalysts for

Liquid-Phase Suzuki Cross-Coupling Reactions

Abstract

We investigate the catalytic activity obtained for three types of nanocatalysts: colloidal supported metal nanoparticles (CSMNs) prepared with silica colloids in solution that are air-dried, CSMNs prepared with dry silica colloids resuspended in doubly deionized water that are air-dried, and palladium nanoparticles loaded onto bulk silica dispersed in doubly deionized water that are air-dried. The three types of catalysts are prepared with and without the aminopropyltriethoxysilane (APTES) linker for a total of six different catalysts that are tested for the liquid-phase crosscoupling reaction between phenylboronic acid and iodobenzene to form biphenyl.

TEM images and EDS spectra were obtained for the six different catalysts to see if the palladium nanoparticles are covalently attached to the functionalized silica colloids and functionalized bulk silica supports and to see if the palladium nanoparticles are adsorbed onto the unfunctionalized silica colloids and unfunctionalized bulk silica supports. The six different catalysts are tested for the Suzuki reaction between phenylboronic acid and iodobenzene to form biphenyl. The CSMNs prepared using unfunctionalized wet silica colloids reacted with palladium nanoparticles that are airdried resulted in the highest catalytic activity. In the case of the nanocatalysts prepared using the APTES linker for covalent attachment of the palladium nanoparticles to the wet silica colloids or bulk silica dispersed in doubly deionized

54

water that are air-dried, there is lower catalytic activity compared to their counterparts prepared without the use of the APTES linker. This suggests that the APTES could be acting as a catalyst poison resulting in lower catalytic activity for the Suzuki reaction between phenylboronic acid and iodobenzene to form biphenyl compared to unfunctionalized CSMNs.

Introduction

Transition metal nanoparticles are very attractive catalysts since they have a large surface-to-volume ratio which makes their surface atoms very active. Their large surface-to-volume ratio results in higher catalytic efficiency per gram than larger bulk catalytic materials. There have been numerous review articles that have focused on the use of homogeneous catalysts (transition metal nanoparticles in colloidal solution) 1-9 , heterogeneous catalysts (transition metal nanoparticles adsorbed onto different supports) 6, 9-20 , and lithographic arrays of nanoparticles 21-25 for many different organic and inorganic reactions. There also have been some studies involving the use of intermediate types of supports such as the use of carbon nanotubes 26-29 and carbon nanofibers 30 as support materials.

Suzuki cross-coupling reactions involve reacting arylboronic acids with aryl halides to form biaryls. There have been many reports on the use of transition metal nanoparticles as catalysts for Suzuki reactions 31-36 . Dendrimer supported palladium nanoparticles have been used as effective nanocatalysts for Suzuki cross-coupling reactions 37, 38 . Shape dependent catalytic activity of copper oxide supported palladium nanoparticles have been tested as nanocatalysts for the Suzuki cross-coupling

55

reactions 36 . Mesoporous silica supported palladium nanoparticles have been used as catalysts for Suzuki reactions 39 . The catalytic activity for the Suzuki cross-coupling reactions have been compared for the use of palladium nanoparticles in solution and those adsorbed onto carbon nanotubes 40 . Palladium nanoparticles have also been used as catalysts for the Suzuki reaction using water as the solvent 41 .

The CSMNs have a high surface area, highly active surface atoms of the nanoparticles, can be suspended in solution during catalysis, and can be easily separated from the reaction mixture. It can be seen that the colloidal supported metal nanoparticles (CSMNs) combine the advantages of heterogeneous catalysts in a near homogeneous format. Also, the ability to be suspended in solution is especially important for liquid-phase reactions. The CSMNs will have higher catalytic activity than homogeneous nanocatalysts (transition metal nanoparticles in colloidal solution) and heterogeneous nanocatalysts (transition metal nanoparticles adsorbed onto bulk supports). Previously we have reported the synthesis of colloidal supported metal nanoparticles (CSMNs) prepared using mercaptopropyl trimethoxysilane (MPTMS) and aminopropyl triethoxysilane (APTES) functionalized silica colloids 42 . MPTMS was chosen as the linker since it has a thiol group that the palladium nanoparticles can bind to and form Pd-S bond. APTES was chosen as a linker since it has a nitrogen group that the palladium nanoparticles can bind to and form the Pd-N bond.

Palladium nanoparticles were covalently attached onto the MPTMS or APTES functionalized silica colloids. The preparation of the CSMNS involves four steps: synthesis of silica colloids, synthesis of palladium nanoparticles, functionalizing the silica colloids with MPTMS and APTES, and covalently attaching the palladium

56

nanoparticles to the functionalized silica colloids. In the case of the CSMNs prepared using MPTMS functionalized silica colloids, there is no catalytic activity during the entire reaction, which was run for 24 hours. MPTMS has a sulfur group which acts as a catalyst poison. The CSMNs prepared using APTES functionalized silica colloids exhibits catalytic activity during the first hour of the reaction.

In this paper, we report the use of different types of CSMNs as catalysts for the

Suzuki cross-coupling reaction between phenylboronic acid and iodobenzene to form biphenyl. The CSMNs that we have investigated consist of metal nanoparticles covalently bound to functionalized silica colloids or adsorbed onto unfunctionalized silica colloids. Out of the six different nanocatalysts, the CSMNs prepared using palladium nanoparticles adsorbed onto unfunctionalized wet silica colloids in solution that are air-dried results in the greatest catalytic activity. In the case of the CSMNs prepared by using palladium nanoparticles covalently attached to the APTES functionalized silica colloids in solution that are air-dried and palladium nanoparticles covalently attached to the APTES functionalized bulk silica dispersed in doubly deionized water that are air-dried, the catalytic activity is smaller than their unfunctionalized counterparts. This suggests that the APTES could be acting as a catalyst poison. Also, the CSMNs prepared by using palladium nanoparticles adsorbed onto unfunctionalized silica colloids in solution that are air-dried resulted in higher catalytic activity than the palladium nanoparticles in colloidal solution as nanocatalysts (containing the same number of grams of palladium).

57

Experimental

Synthesis of Silica Colloids

The silica colloids were synthesized by using the Stoeber synthesis method 43 .

First, 30 mL of ethanol and 2.4 mL of ammonium hydroxide was added to an

Erlenmeyer flask and stirred for 5 minutes. Then, 1.2 mL of tetraethylorthosilicate

(TEOS) was added to the solution containing ethanol and ammonium hydroxide and this solution was stirred overnight. The solution was initially clear and after 15-20 minutes, the solution starts to turn cloudy and the final solution is very turbid and consists of a suspension of the silica colloids.

Synthesis of PVP Stabilized Palladium Nanoparticles

The palladium nanoparticles were synthesized by using the ethanol reduction method similar to that described previously 44-47 . The palladium precursor solution

(H

2

PdCl

4

) was prepared by adding 0.0887 g of PdCl

2

and 6 mL of 0.2 M HCl, and diluting to 250 mL with doubly deionized water. A solution containing 15 mL of 2 mM of H

2

PdCl

4

, 21 mL of doubly deionized water, 0.0667 g of PVP, and 4 drops of 1

M HCl was heated. When the solution began to reflux, 14 mL of ethanol was added.

The solution was then refluxed for 3 h, and this resulted in a dark brown suspension of

Pd nanoparticles.

Functionalization of Silica Colloids

We have used 3-aminopropyl triethoxysilane (APTES) as the linker to functionalize the silica colloids. The functionalized silica colloids were prepared in

58

which 100 " L of APTES is added to the silica colloid suspension and allowed to react for 6 hours. We have also prepared unfunctionalized silica colloids for direct adsorption of the metal nanoparticles onto the surface of the unfunctionalized silica colloids.

Synthesis of Colloidal Supported Palladium Nanoparticles (CSMNs) Using

Functionalized and Unfunctionalized Silica Colloid Suspensions

In the case of the functionalized wet silica colloids, 100 mL of 3aminopropyltriethoxysilane (APTES) was added and allowed to react for six hours.

Two mL of the centrifuged functionalized or unfunctionalized silica colloids and four mL of the PVP capped palladium nanoparticles are placed into a scintillation vial and this suspension is mixed for 24 hours to allow the palladium nanoparticles to covalently bind to the functionalized silica colloids or adsorb onto the unfunctionalized silica colloids. The CSMNs prepared with palladium nanoparticles covalently attached to the functionalized silica colloids or adsorbed onto the unfunctionalized silica colloids were then centrifuged four times at 13,500 rpm for 3 minutes each time. During the first two centrifugation cycles, the CSMNs were redispersed in ethanol and during the last two centrifugation cycles, they were redispersed in doubly deionized water. After the final centrifugation cycle, the supernatant is removed and the CSMNs prepared with functionalized and unfunctionalized silica colloids are allowed to air-dry, and 0.1 g of each type of

CSMNs is used as catalysts for the Suzuki cross-coupling reaction between phenylboronic acid and iodobenzene.

59

Synthesis of Colloidal Supported Metal Nanoparticles (CSMNs) Using Functionalized and Unfunctionalized Dry Silica Colloids Resuspended in Doubly Deionized Water

The silica colloid suspensions were synthesized by the Stoeber synthesis method 42 . The silica colloids were first centrifuged four times at 13,500 rpm for 3 minutes each time. During the first two centrifugation cycles, the functionalized and unfunctionalized silica colloids were redispersed in ethanol and during the last two centrifugation cycles, they were redispersed in doubly deionized water. After the final cycle of centrifugation, the supernatant was removed and the silica colloids were allowed to air-dry overnight.

In the case of the functionalized silica colloids, 0.32 g of the dried silica colloids is mixed with 30 mL of doubly deionized water and stirred overnight. To this suspension of the initial dried colloids, 100 mL of APTES is added to the silica colloid suspension and allowed to react for 6 hours. In the case of the unfunctionalized silica colloids, 0.32 g of the dried silica colloids is mixed with 30 mL of doubly deionized water and stirred overnight. Two mL of the centrifuged functionalized or unfunctionalized silica colloids and four mL of the PVP capped palladium nanoparticles are placed into a scintillation vial and this suspension is mixed for 24 hours to allow the palladium nanoparticles to covalently bind to the functionalized silica colloids or adsorb onto the unfunctionalized silica colloids. The CSMNs were then centrifuged four times at 13,500 rpm for 3 minutes each time. During the first two centrifugation cycles, the CSMNs prepared using functionalized and unfunctionalized silica colloids were redispersed in ethanol and during the last two centrifugation cycles, they were redispersed in doubly deionized water. After the final

60

centrifugation cycle, the CSMNs are allowed to air-dry overnight. Then, 0.1 gram of the two types of CSMNs is used as catalysts for the Suzuki cross-coupling reaction between phenylboronic acid and iodobenzene to form biphenyl.

Preparation of Palladium Nanoparticles Loaded onto Functionalized and

Unfunctionalized Bulk Silica Dispersed in Doubly Deionized Water

In the case of the functionalized and unfunctionalized bulk silica, 0.32 g of the bulk dry silica and 30 mL of doubly deionized water were added to an Erlenmeyer flask and stirred overnight. In the case of the functionalized bulk silica, 100 mL of 3aminopropyl triethoxysilane was added and reacted for 6 hours. Two mL of the centrifuged functionalized or unfunctionalized bulk silica and four mL of the PVP capped palladium nanoparticles are placed into a scintillation vial and this suspension is mixed for 24 hours to allow the palladium nanoparticles to bind to the functionalized bulk silica or adsorb onto the unfunctionalized bulk silica. The bulk silica was then centrifuged four times at 13,500 rpm for 3 minutes each time. During the first two centrifugation cycles, the functionalized and unfunctionalized bulk silica loaded with palladium nanoparticles were redispersed in ethanol and during the last two centrifugation cycles, they were redispersed in doubly deionized water. After the final cycle of centrifugation, the supernatant is removed and the functionalized and unfunctionalized bulk silica loaded with palladium nanoparticles are allowed to air-dry overnight. Then, 0.1 g of palladium nanoparticles loaded onto functionalized and unfunctionalized bulk silica are used as the catalysts for the Suzuki cross-coupling reaction between phenylboronic acid and iodobenzene to form biphenyl.

61

Characterization Studies with TEM and EDS

One drop of dilute suspensions of colloidal supported palladium nanoparticles and palladium nanoparticles loaded onto bulk silica was placed on Formvar coated copper grids. The drop was allowed to air-dry for ~1 hour. The JEOL 2100EX TEM was used to obtain the TEM images of the colloidal supported metal nanoparticles

(CSMNs) and palladium nanoparticles covalently attached or adsorbed onto silica nanoparticles. Energy dispersive spectroscopy (EDS) was used to determine what elements are present in the CSMNs.

Suzuki Cross-Coupling Reaction

The Suzuki reaction between phenylboronic acid and iodobenzene was catalyzed using the PVP-Pd nanoparticles as described previously 44, 48, 49 . For this reaction, 0.49 g (6 mmol) of sodium acetate, 0.37 g (3 mmol) of phenylboronic acid, and 0.20 g (1 mmol) of iodobenzene was added to 150 mL of 3:1 acetonitrile:water solvent. The solution was heated to 100 o C and 0.1 gram of the CSMNs or palladium nanoparticles loaded onto bulk silica was added to start the reaction. The reaction mixture was refluxed for a total of 12 hours. Kinetic studies were conducted during the first hour of the reaction.

Results and Discussion

The CSMNs have a high surface area, highly active surface atoms of the nanoparticles, can be suspended in solution during catalysis, and can be easily separated from the reaction mixture. The colloidal supported metal nanoparticles

62

(CSMNs) combine the advantages of heterogeneous catalysts in a near homogeneous format. The ability of the CSMNs to be suspended in solution is especially important for liquid-phase reactions. The CSMNs will have higher catalytic activity than homogeneous nanocatalysts (transition metal nanoparticles in colloidal solution) and heterogeneous nanocatalysts (transition metal nanoparticles adsorbed onto bulk supports). We discuss the synthesis and characterization of different types of colloidal supported metal nanoparticles (CSMNs) and palladium nanoparticles loaded onto bulk silica as catalysts for the Suzuki cross-coupling reaction between phenylboronic acid and iodobenzene. We also discuss the results of the kinetics experiments using HPLC for the Suzuki cross-coupling reaction between phenylboronic acid and iodobenzene to form biphenyl for the different types of nanocatalysts.

Synthesis of Different Types of CSMNs and Palladium Nanoparticles Loaded onto

Bulk Silica

Six different types of nanocatalysts were synthesized as catalysts for the liquidphase Suzuki cross-coupling reaction. The palladium nanoparticles have an average size of 2.9 nm as reported previously 42 . In all cases, the palladium nanoparticles are adsorbed or covalently attached to the surface of the silica colloids. The silica colloids are not hollow or porous so the palladium nanoparticles cannot be adsorbed or covalently attached inside the silica colloids. Two types of CSMNs involve the functionalized and unfunctionalized wet silica colloids reacting with the palladium nanoparticles to form CSMNs. In these two types of CSMNs, the term wet silica colloid refers to the silica colloids being in aqueous solution. Preparing the CSMNs

63

by adsorption of the palladium nanoparticles directly to the silica colloid suspensions involves fewer steps compared to covalently attaching the palladium nanoparticles to

APTES functionalized silica colloid suspensions. Using the wet silica colloids involves adsorption or covalent attachment of the palladium nanoparticles to the silica colloids in its native form. Figure 1 shows a representative TEM image and EDS spectrum for the CSMNs prepared using aminopropyl triethoxysilane (APTES) functionalized suspended wet silica colloids. In this case, the palladium nanoparticles are covalently attached to the functionalized silica colloids. Figure 2 shows a representative TEM image and EDS spectrum for the CSMNs prepared using unfunctionalized suspended wet silica colloids. In this case, the palladium nanoparticles are adsorbed onto the unfunctionalized wet silica colloids. The black specks in the TEM images are the palladium nanoparticles covalently attached to the functionalized silica colloid or adsorbed onto unfunctionalized silica colloids. In the

EDS spectra for both cases, the Pd and Si peaks are present confirming that the Pd nanoparticles are loaded onto the silica colloids. The copper peaks that are present are due to the copper TEM grids. These two CSMNs that are formed are allowed to airdry and are used as catalysts for the Suzuki cross-coupling reaction between phenylboronic acid and iodobenzene, which will be discussed later in the paper.

Another two types of CSMNs involve functionalized and unfunctionalized dried silica colloids resuspended in doubly deionized water and reacting with the palladium nanoparticles to form CSMNs. In these two cases, the silica colloids are air-dried and then resuspended in doubly deionized water. Preparing the CSMNs by adsorption of the palladium nanoparticles directly to the silica colloid suspensions

64

involves fewer steps compared to covalently attaching the palladium nanoparticles to

APTES functionalized silica colloid suspensions. Using the dry silica colloids resuspended in doubly deionized water for adsorption or covalent attachment of the palladium nanoparticles is not in its native form, which is a disadvantage of this method compared to using the wet silica colloids in their native form. Figure 3 shows a representative TEM image and EDS spectrum for the CSMNs prepared using aminopropyl triethoxysilane (APTES) functionalized dried silica colloids resuspended in doubly deionized water. In this case, the palladium nanoparticles are covalently attached to the functionalized dried silica colloids resuspended in doubly deionized water. Figure 4 shows a representative TEM image and EDS spectrum for the CSMNs prepared using unfunctionalized dried silica colloids resuspended in doubly deionized water. In this case, the palladium nanoparticles adsorb onto the surface of the dried silica colloids resuspended in doubly deionized water. The black specks in the TEM images are the palladium nanoparticles covalently attached to the functionalized silica colloids in Figure 3 or adsorbed onto unfunctionalized silica colloids in Figure 4. In the EDS spectra for both cases, the Pd and Si peaks are present confirming that the Pd nanoparticles are loaded onto the Si colloids. The copper peaks that are present are due to the copper TEM grids. These two types of CSMNs are then air-dried and used as catalysts for the Suzuki cross-coupling reaction which is discussed later on in this paper.

In addition, two types of palladium nanoparticles covalently attached or adsorbed onto functionalized and unfunctionalized dry bulk silica dispersed in doubly deionized water were prepared. In these two cases, the bulk silica supports are

65

dispersed in doubly deionized water. Adsorption of the palladium nanoparticles directly to the bulk silica dispersed in water involves fewer steps compared to covalently attaching the palladium nanoparticles to APTES functionalized bulk silica dispersed in doubly deionized water. Figure 5 shows a representative TEM image and

EDS spectrum for palladium nanoparticles supported on APTES functionalized dried bulk silica dispersed in doubly deionized water. In this case, the palladium nanoparticles are covalently attached onto the functionalized bulk silica. Figure 6 shows a representative TEM image and EDS spectrum for palladium nanoparticles supported on unfunctionalized dried bulk silica dispersed in water. In this case, the palladium nanoparticles are adsorbed onto the unfunctionalized bulk silica. In the

EDS spectra for both cases, the Pd and Si peaks are present confirming that the Pd

NPs are loaded onto the bulk silica. The palladium nanoparticles covalently attached or adsorbed onto bulk silica are air-dried and used as catalysts for the Suzuki crosscoupling reaction between phenylboronic acid and iodobenzene as discussed later in the paper.

In the case of the nanocatalysts consisting of the palladium nanoparticles adsorbing onto the unfunctionalized silica colloids, the free sites on the palladium nanoparticles (sites not occupied by the PVP stabilizer) are accessible to act as a catalyst. Also, during the adsorption process, the PVP on the nanoparticle surface can be displaced and this results in more sites that can act as catalysts. In the case of the nanocatalysts consisting of the palladium nanoparticles covalently attached to the

APTES functionalized silica colloids, the free sites on the palladium nanoparticles

(sites not occupied by the PVP stabilizer) are accessible to act as a catalyst. In

66

addition, it is possible that during the process of the palladium nanoparticles covalently binding to the APTES functionalized silica colloids, some of the PVP could get displaced. This would result in more free sites available to act as catalysts.

Nanocatalysts for Liquid-Phase Suzuki Cross-Coupling Reactions

We have tested the use of the different types of nanocatalysts for the Suzuki cross-coupling reaction between phenylboronic acid and iodobenzene to form biphenyl. The Pd content of all of the nanocatalysts is the same since we prepared them in the same way. We allowed the silica colloids and bulk silica to react with 5 mL of the palladium nanoparticles and centrifuged them. The supernatant is clear indicating that all of the palladium nanoparticles have been adsorbed or covalently bound to the silica colloids and bulk silica. We have also compared the catalytic activity of these catalysts to that of 5 mL of the colloidal palladium nanoparticles as catalysts. We have used reversed phase HPLC to follow the kinetics during the first hour of the Suzuki reaction for the different nanocatalysts we have prepared such as wet functionalized and unfunctionalized silica colloids in solution reacted with palladium nanoparticles, dried functionalized and unfunctionalized silica colloids resuspended in water reacted with palladium nanoparticles, dry functionalized and unfunctionalized bulk silica dispersed in doubly deionized water reacted with palladium nanoparticles, and palladium nanoparticles in colloidal solution. In the six cases involving CSMNs or palladium nanoparticles loaded onto bulk silica, the nanocatalysts are allowed to air-dry and 0.1 g of each catalyst were used for the

Suzuki cross-coupling reaction between phenylboronic acid and iodobenzene. We

67

have made comparisons of the catalytic activity for these different types of nanocatalysts.

We have compared the catalytic activity for CSMNs consisting of palladium nanoparticles covalently attached to APTES functionalized wet silica colloids to those of CSMNs consisting of palladium nanoparticles covalently attached to the APTES functionalized dry silica colloids resuspended in doubly deionized water, and palladium nanoparticles covalently attached to the APTES functionalized dry bulk silica dispersed in doubly deionized water. In all three cases, the nanocatalysts are allowed to air-dry and 0.1 gram of each catalyst were used for the Suzuki crosscoupling reaction between phenylboronic acid and iodobenzene. Figure 7 compares the catalytic activity of these three types of nanocatalysts based on the HPLC kinetics study during the first hour of the reaction. The catalytic activity of the CSMNs consisting of palladium nanoparticles covalently attached to the APTES functionalized wet silica colloids that are air-dried is the highest for the three types of nanocatalysts.

The CSMNs consisting of palladium nanoparticles covalently attached to the APTES functionalized dry silica colloids resuspended in doubly deionized water that are airdried results in an intermediate catalytic activity. The palladium nanoparticles covalently attached to the APTES functionalized dry bulk silica dispersed in doubly deionized water that are air-dried results in the lowest catalytic activity out of the three types of nanocatalysts involving the palladium nanoparticles being covalently attached to the APTES functionalized silica colloids or bulk silica.

We have compared the catalytic activity of CSMNs consisting of palladium nanoparticles adsorbed onto wet silica colloids to those of CSMNs consisting of

68

palladium nanoparticles adsorbed onto the dry silica colloids resuspended in doubly deionized water, and palladium nanoparticles adsorbed onto the dry bulk silica dispersed in doubly deionized water. In all three cases, the nanocatalysts are allowed to air-dry and 0.1 gram of each catalyst were used for the Suzuki cross-coupling reaction between phenylboronic acid and iodobenzene. Figure 8 compares the catalytic activity of these three types of nanocatalysts based on the HPLC kinetics study during the first hour of the reaction. The CSMNs consisting of palladium nanoparticles adsorbed onto the wet silica colloids that are air-dried results in the highest catalytic activity. In this case, biphenyl is formed as soon as the nanocatalyst is added. In the case of the CSMNs consisting of palladium nanoparticles adsorbed onto the dry silica colloids resuspended in doubly deionized water that are air-dried and the palladium nanoparticles adsorbed onto the dry bulk silica redispersed in doubly deionized water that are air-dried, the catalytic activity is much lower than that of CSMNs consisting of palladium nanoparticles adsorbed onto the wet silica colloids that are air-dried.

The CSMNs consisting of palladium nanoparticles adsorbed onto the wet silica colloids that are air-dried as shown in Figure 8 results in higher catalytic activity compared to that of the CSMNs consisting of palladium nanoparticles covalently attached to the APTES functionalized wet silica colloids that are air-dried as shown in

Figure 7. This result suggests that the APTES could be poisoning the catalytic activity for the Suzuki cross-coupling reaction between phenylboronic acid and iodobenzene.

Previously we have observed that if we use palladium nanoparticles covalently attached to mercaptopropyl trimethoxysilane (MPTMS) functionalized silica colloids,

69

there is no catalytic activity during the first hour of the reaction. In this case, the mercapto group in the linker results in poisoning of the catalyst. Nitrogen compounds result in poisoning of different catalysts 50-52 . This supports the idea that in the case of the CSMNs consisting of palladium nanoparticles covalently attached to APTES functionalized silica colloids, there could be some poisoning of the catalyst due to

APTES, which contain an amine group which results in the lower catalytic activity that is observed.

We have also compared the catalytic activity of the CSMNs consisting of palladium nanoparticles adsorbed onto unfunctionalized wet silica colloids that are airdried to that of palladium nanoparticles in colloidal solution. Both of these catalysts have the same number of grams of Pd. As can be seen in Figure 9, the CSMNs consisting of palladium nanoparticles adsorbed onto the unfunctionalized wet silica colloids that are air-dried results in larger catalytic activity compared to the palladium nanoparticles in colloidal solution. In addition to the higher catalytic activity of the

CSMNs consisting of palladium nanoparticles adsorbed onto unfunctionalized wet silica colloids, these CSMNs have additional advantages such as having high surface area and being easily separated from the reaction mixture. This makes using the

CSMNs consisting of palladium nanoparticles adsorbed onto the unfunctionalized wet silica colloids very attractive to use compared to the palladium nanoparticles in colloidal solution.

We have also compared the catalytic activity of the CSMNs consisting of palladium nanoparticles adsorbed to the wet silica colloid surface that are air-dried to those of the CSMNs consisting of palladium nanoparticles covalently attached to the

70

APTES functionalized wet silica colloids that are air-dried as can be seen in Figure 10.

In all three cases, the nanocatalysts are allowed to air-dry and 0.1 g of each catalyst was used for the Suzuki cross-coupling reaction between phenylboronic acid and iodobenzene. The CSMNs consisting of palladium nanoparticles adsorbed onto the wet silica colloids that are air-dried results in higher catalytic activity compared to the

CSMNs consisting of palladium nanoparticles covalently attached to the APTES functionalized wet silica colloids that are air-dried. The TEM images of the two types of nanocatalysts are similar, yet there are differences in the catalytic activity. The lower catalytic activity for the CSMNs consisting of palladium nanoparticles covalently attached to the APTES functionalized wet silica colloids that are air-dried could be due to APTES poisoning the catalyst resulting in lower catalytic activity.

This poisoning effect could explain why there is lower catalytic activity in the case of the palladium nanoparticles being covalently attached to the APTES functionalized silica colloids in spite of the similar TEM images.

As can be seen in Figure 11, we have compared the catalytic activity of the

CSMNs consisting of palladium nanoparticles adsorbed onto dried silica colloids resuspended in doubly deionized water that are air-dried to that of the CSMNs consisting of palladium nanoparticles covalently attached to the APTES functionalized dried silica colloids resuspended in doubly deionized water that are air-dried. In both cases, the nanocatalysts are allowed to air-dry and 0.1 g of each catalyst were used for the Suzuki cross-coupling reaction between phenylboronic acid and iodobenzene.

There is very little difference in the catalytic activity for the two types of nanocatalysts for the Suzuki cross-coupling reaction between phenylboronic acid and iodobenzene

71

to form biphenyl. As shown in Figure 8, the CSMNs consisting of palladium nanoparticles adsorbed onto unfunctionalized wet silica colloids that are air-dried have much higher catalytic activity than the CSMNs consisting of palladium nanoparticles adsorbed onto unfunctionalized dried silica colloids resuspended in doubly deionized water that are air-dried as shown in Figure 11 and CSMNs consisting of palladium nanoparticles covalently attached to the APTES functionalized dried silica colloids resuspended in doubly deionized water that are air-dried as shown in Figure 11.

We have also compared the catalytic activity of the palladium nanoparticles adsorbed onto dry bulk silica dispersed in doubly deionized water that are air-dried to that of the palladium nanoparticles covalently attached onto APTES functionalized dry bulk silica dispersed in doubly deionized water that are air-dried. In all three cases, the nanocatalysts are allowed to air-dry and 0.1 g of each catalyst was used for the

Suzuki reaction between phenylboronic acid and iodobenzene. As shown in Figure

12, the palladium nanoparticles adsorbed onto dry bulk silica dispersed in doubly deionized water that are air-dried results in higher catalytic activity. It can also be seen that the palladium nanoparticles covalently attached onto APTES functionalized dry bulk silica dispersed in doubly deionized water that are air-dried results in lower catalytic activity. The palladium nanoparticles covalently attached to bulk silica dispersed in doubly deionized water functionalized with 200 mL APTES that are airdried resulted in lower catalytic activity than palladium nanoparticles covalently attached to bulk silica dispersed in doubly deionized water functionalized with 100 mL APTES that are air-dried. This trend supports the idea that the APTES acts as a catalyst poison resulting in lower catalytic activity.

72

Leaching Test

In the case of transition metal nanoparticles, there have been recent studies in which it has been found that heterogeneous catalysts are not truly heterogeneous in which the catalyst leaches into solution 53, 54 . This raises the question about whether catalysis is homogeneous or heterogeneous. Our CSMNs combine the advantages of heterogeneous catalysts in a near homogeneous format. We have conducted experiments to determine if leaching of the palladium atoms into solution occurs. We have refluxed the CSMNs in 3:1 acetonitrile:water mixture for one hour. Then the solution is centrifuged to remove the CSMNs in which leached atoms would be in the supernatant solution. We used the supernatant solution as catalysts for the Suzuki reaction and compared the catalytic activity to that of the CSMNs catalyzing the

Suzuki reaction. The results are shown in Figure 13 and it can be seen that while there is some catalytic activity in the case of the supernatant solution, there is a much greater amount of catalytic activity in the case of the CSMNs used as catalysts.

Overall, while there is a small amount of leaching that occurs, the overall catalytic activity is much greater when the CSMNs are used as catalysts.

Conclusions

The catalytic activity is obtained for three types of nanocatalysts: colloidal supported metal nanoparticles (CSMNs) prepared with silica colloids in solution that are air-dried, CSMNs prepared with dry silica colloids that is resuspended in doubly deionized water that are air-dried, and palladium nanoparticles loaded onto bulk silica dispersed in doubly deionized water that are air-dried. These three types of catalysts

73

are prepared with and without the aminopropyl triethoxysilane (APTES) linker for a total of six different catalysts that are tested for the liquid-phase cross-coupling reaction between phenylboronic acid and iodobenzene to form biphenyl. The CSMNs prepared with the silica colloids in solution without the use of the APTES linker that are air-dried results in the highest catalytic activity for the Suzuki cross-coupling reaction. In the case of the CSMNs consisting of palladium nanoparticles covalently bound to APTES functionalized wet silica colloids that are air-dried and palladium nanoparticles covalently bound to the APTES functionalized dry bulk silica dispersed in doubly deionized water that are air-dried, it is observed that there is lower catalytic activity than their counterparts prepared without the use of the APTES linker. This suggests that the APTES could be acting as a catalyst poison resulting in lower catalytic activity for the Suzuki reaction. Also, it was observed that the CSMNs prepared with the silica colloids in solution without the use of the APTES linker that are air-dried results in higher catalytic activity than the use of palladium nanoparticles in colloidal solution containing the same number of grams of palladium. Based on these results, it can be seen that the CSMNs prepared using the palladium nanoparticles adsorbed onto unfunctionalized silica colloids in solution that are airdried result in the highest catalytic activity for the Suzuki reaction between phenylboronic acid and iodobenzene to form biphenyl.

Acknowledgments

74

We thank Dr. Richard Kingsley at the University of Rhode Island Electron

Microscopy Center for assistance in obtaining the TEM images. We thank the

University of Rhode Island for the start-up funds for funding of this work.

75

Figure 3-1— Example TEM image showing the palladium nanoparticles covalently bound to amine functionalized silica colloid suspensions (a) and EDS spectrum, which reflects the presence of both Pd and Si (b). The Cu peaks are due to the copper TEM grids.

76

Figure 3-2— Examples of TEM images showing palladium nanoparticles directly adsorbed onto the surface of the silica colloid suspensions (a) and EDS spectrum, which reflects the presence of both Pd and Si (b). The Cu peaks are due to Cu TEM grids.

77

Figure 3-3— A representative TEM image showing the palladium nanoparticles covalently bound to APTES functionalized dried silica colloid suspensions (a) and

EDS spectrum, which reflects the presence of both Pd and Si (b). The Cu peaks are due to the copper TEM grids.

78

Figure 3-4— A representative TEM image showing the palladium nanoparticles adsorbed onto the unfunctionalized dried silica colloid suspensions (a) and EDS spectrum, which reflects the presence of both Pd and Si (b). The Cu peaks are due to the copper TEM grids.

79

Figure 3-5— Example of TEM image showing APTES functionalized bulk silica with palladium nanoparticles covalently attached to its surface (a) and EDS spectrum which reflects the presence of both Pd and Si (c). The Cu peaks are due to the copper TEM grids.

80

Figure 3-6— Example of TEM image showing unfunctionalized bulk silica with palladium nanoparticles adsorbed to its surface (a) and EDS spectrum which reflects the presence of both Pd and Si (b). The Cu peaks that appear are due to copper TEM grids.

81

2.5

2.0

1.5

Wet Si Colloids Reacted with Pd NPs with 100 µ L amine linker

Dry Si Colloids Reacted with Pd NPs with 100 µ L amine linker

Bulk Dry Si Reacted with Pd NPs with 100 µ L amine linker

1.0

0.5

0.0

0 10 20 30 40

Time (min.)

50 60

Figure 3-7— Comparisons of the catalytic activity with wet APTES functionalized Si colloids reacted with Pd nanoparticles that are air-dried, dry APTES functionalized Si colloids resuspended in doubly deionized water and reacted with Pd nanoparticles that are air-dried, and bulk dry APTES functionalized Si dispersed in doubly deionized water and reacted with Pd nanoparticles that are air-dried.

3.0

Wet Si Colloids Reacted with Pd NPs without linker

Dry Si Colloids Reacted with Pd NPs without linker

Bulk Dry Si Reacted with Pd NPs without linker

2.5

2.0

1.5

1.0

0.5

0.0

0 10 20 30 40

Time (min.)

50 60

Figure 3-8— Comparisons of the catalytic activity with wet unfunctionalized Si colloids reacted with Pd NPs that are air-dried, dry unfunctionalized Si colloids resuspended in doubly deionized water and reacted with Pd NPs that are air-dried, and unfunctionalized bulk dry Si dispersed in doubly deionized water and reacted with Pd

NPs that are air-dried.

82

3.0

2.5

2.0

1.5

1.0

Wet Si Colloids Reacted with Pd NPs WIthout Linker

5 mL of Palladium Nanoparticles

0.5

0.0

0 10 20 30 40 50

Time (min.)

60

Figure 3-9— Comparisons of the catalytic activity with wet Si colloids reacted with

Pd NPs without linker that are air-dried and palladium nanoparticles in colloidal solution.

3.0

2.5

2.0

1.5

1.0

Wet Si Colloids Reacted with Pd NPs Without Linker

Wet Si Colloids Reacted with Pd NPs Using 100 µ L Linker

Wet Si Colloids Reacted with Pd NPs Using 200 µ L Linker

0.5

0.0

0 10 20 30 40

Time (min.)

50 60

Figure 3-10— Comparisons of the catalytic activity with wet unfunctionalized Si colloids reacted with Pd nanoparticles that are air-dried, and two different wet APTES functionalized Si colloids reacted with Pd nanoparticles that are air-dried

83

1.6

1.4

1.2

1.0

0.8

0.6

0.4

0.2

0.0

0

Dried Si Colloids Reacted with Pd NPs Without Linker

Dried Si Colloids Reacted with Pd NPs With 100 µ L Amine Linker

10 20 30 40

Time (min.)

50 60

Figure 3-11— Comparisons of the catalytic activity with unfunctionalized dried Si colloids resuspended in doubly deionized water that are air-dried and APTES functionalized dried Si colloids resuspended in doubly deionized water that are airdried

1.6

1.4

1.2

1.0

0.8

0.6

0.4

0.2

0.0

0

Bulk Si Reacted with Pd NPs Without Linker

Bulk Si Reacted with Pd NPs Using 100 µ L Amine Linker

Bulk Si Reacted with Pd NPs Using 200 µ L Amine Linker

10 20 30 40

TIme (min)

50 60

Figure 3-12— Comparisons of the catalytic activity with unfunctionalized bulk dry Si dispersed in doubly deionized water that are air-dried and APTES functionalized bulk dry Si dispersed in doubly deionized water that are air-dried

84

3.0

2.5

2.0

1.5

1.0

0.5

0.1g of CSMNs

Supernatant - CSMNs separated

0.0

0 10 20 30 40 50 60

Time (min.)

Figure 3-13 —Comparison of the catalytic activity of the supernatant solution after refluxing the CSMNs in 3:1 acetonitrile and centrifuging to that of the CSMNs used as catalysts for the Suzuki reaction.

85

References

1) Astruc, D., Palladium catalysis using dendrimers: molecular catalysts versus nanoparticles. Tetrahedron: Asymmetry 21, (9-10), 1041-1054

2) Boennemann, H.; Nagabhushana, K. S., Colloidal nanoparticles in catalysis.

Surface and Nanomolecular Catalysis 2006 , 63-94

3) Boennemann, H.; Nagabhushana, K. S.; Richards, R. M., Colloidal nanoparticles stabilized by surfactants or organo-aluminum derivatives: preparation and use as catalyst precursors. Nanoparticles and Catalysis 2008 ,

49-91.

4) Brock, S. L.; Senevirathne, K., Recent developments in synthetic approaches to transition metal phosphide nanoparticles for magnetic and catalytic applications. Journal of Solid State Chemistry 2008, 181, (7), 1552-1559.

5) Bronstein, L. M.; Matveeva, V. G.; Sulman, E. M., Nanoparticulate catalysts based on nanostructured polymers. Nanoparticles and Catalysis 2008 , 93-127.

6) Somorjai, G. A.; Contreras, A. M.; Montano, M.; Rioux, R. M., Clusters, surfaces, and catalysis. Proceedings of the National Academy of Sciences of the

United States of America 2006, 103, (28), 10577-10583.

7) Duran Pachon, L.; Rothenberg, G., Transition-metal nanoparticles: synthesis, stability and the leaching issue. Applied Organometallic Chemistry 2008, 22,

(6), 288-299.

8) Gates, B. C., Individual nanoparticles in action. Nature Nanotechnology 2008,

3, (10), 583-584.

9) Gorodetskii, V. V.; Sametova, A. A.; Matveev, A. V.; Tapilin, V. M., From single crystals to supported nanoparticles in experimental and theoretical studies of H2 oxidation over platinum metals (Pt, Pd): Intermediates, surface waves and spillover. Catalysis Today 2009, 144, (3-4), 219-234.

10) Arenz, M.; Landman, U.; Heiz, U., CO combustion on supported gold clusters.

ChemPhysChem 2006, 7, (9), 1871-1879.

86

11) Barkhuizen, D.; Mabaso, I.; Viljoen, E.; Welker, C.; Claeys, M.; van Steen, E.;

Fletcher, J. C. Q., Experimental approaches to the preparation of supported metal nanoparticles. Pure and Applied Chemistry 2006, 78, (9), 1759-1769.

12) Budarin, V.; Clark, J. H.; Luque, R.; White, R., Starbons: cooking up nanostructured mesoporous materials. Material Matters (Milwaukee, WI,

United States) 2009, 4, (1), 19-22.

13) Campelo, J. M.; Luna, D.; Luque, R.; Marinas, J. M.; Romero, A. A.,

Sustainable preparation of supported metal nanoparticles and their applications in catalysis. ChemSusChem 2009, 2, (1), 18-45.

14) Somorjai, G. A.; Kliewer, C. J., Reaction selectivity in heterogeneous catalysis: An invited review. Reaction Kinetics and Catalysis Letters 2009, 96,

(2), 191-208.

15) Thomas, J. M.; Raja, R., The advantages and future potential of single-site heterogeneous catalysts. Topics in Catalysis 2006, 40, (1-4), 3-17.

16) Corma, A.; Garcia, H., Supported gold nanoparticles as catalysts for organic reactions. Chemical Society Reviews 2008, 37, (9), 2096-2126.

17) Cuenya, B. R., Synthesis and catalytic properties of metal nanoparticles: size, shape, support, composition, and oxidation state effects. Thin Solid Films 518,

(12), 3127-3150.

18) Fan, J.; Gao, Y., Nanoparticle-supported catalysts and catalytic reactions - a mini-review. Journal of Experimental Nanoscience 2006, 1, (1-4), 457-475.

19) Fukuoka, A.; Dhepe, P. L., Sustainable green catalysis by supported metal nanoparticles. Chemical Record 2009, 9, (4), 224-235.

20) Molenbroek, A. M.; Helveg, S.; Topsoe, H.; Clausen, B. S., Nano-Particles in

Heterogeneous Catalysis. Topics in Catalysis 2009, 52, (10), 1303-1311.

21) Curri, M. L.; Comparelli, R.; Striccoli, M.; Agostiano, A., Emerging methods for fabricating functional structures by patterning and assembling engineered nanocrystals. Physical Chemistry Chemical Physics 12, (37), 11197-11207.

87

22) Eppler, A.; Rupprechter, G.; Guczi, L.; Somorjai, G. A., Model Catalysts

Fabricated Using Electron Beam Lithography and Pulsed Laser Deposition.

Journal of Physical Chemistry B 1997, 101, (48), 9973-9977.

23) Kasemo, B.; Johansson, S.; Persson, H.; Thormahlen, P.; Zhdanov, V. P.,

Catalysis in the nm-regime: manufacturing of supported model catalysts and theoretical studies of the reaction kinetics. Topics in Catalysis 2000, 13, (1,2),

43-53.

24) Somorjai, G. A., New model catalysts (platinum nanoparticles) and new techniques (SFG and STM) for studies of reaction intermediates and surface restructuring at high pressures during catalytic reactions. Applied Surface

Science 1997, 121/122, 1-19.

25) Somorjai, G. A.; York, R. L.; Butcher, D.; Park, J. Y., The evolution of model catalytic systems; studies of structure, bonding and dynamics from single crystal metal surfaces to nanoparticles, and from low pressure (<10-3 Torr) to high pressure (>10-3 Torr) to liquid interfaces. Physical Chemistry Chemical

Physics 2007, 9, (27), 3500-3513.

26) Bottini, M.; Dawson, M. I.; Mustelin, T., Carbon nanotube-based inorganic composites. Chemistry of Carbon Nanotubes 2008, 3, 25-34.

27) Pan, X.; Bao, X., Reactions over catalysts confined in carbon nanotubes.

Chemical Communications (Cambridge, United Kingdom) 2008 , (47), 6271-

6281.

28) Chen, X.; Hou, Y.; Wang, H.; Cao, Y.; He, J., Facile Deposition of Pd

Nanoparticles on Carbon Nanotube Microparticles and Their Catalytic Activity for Suzuki Coupling Reactions. Journal of Physical Chemistry C 2008, 112,

(22), 8172-8176.

29) Karousis, N.; Tsotsou, G.-E.; Evangelista, F.; Rudolf, P.; Ragoussis, N.;

Tagmatarchis, N., Carbon Nanotubes Decorated with Palladium Nanoparticles:

Synthesis, Characterization, and Catalytic Activity. Journal of Physical

Chemistry C 2008, 112, (35), 13463-13469.

30) Bitter, J. H., Carbon nanofibers in catalysis - fundamental studies and scope of application. Nanocatalysis 2006 , 99-125.

88

31) Tsvelikhovsky, D.; Popov, I.; Gutkin, V.; Rozin, A.; Shvartsman, A.; Blum, J.,

On the involvement of palladium nanoparticles in the Heck and Suzuki reactions. European Journal of Organic Chemistry 2009 , (1), 98-102.

32) Ren, L. Z.; Meng, L. J., Suzuki coupling reactions catalyzed by poly(N-ethyl-

4-vinylpyridinium) bromide stabilized palladium nanoparticles in aqueous solution. eXPRESS Polymer Letters 2008, 2, (4), 251-255.

33) Narayanan, R., Recent advances in noble metal nanocatalysts for Suzuki and

Heck cross-coupling reactions. Molecules 15, 2124-2138.

34) Mejias, N.; Pleixats, R.; Shafir, A.; Medio-Simon, M.; Asensio, G., Water-

Soluble Palladium Nanoparticles: Click Synthesis and Applications as a

Recyclable Catalyst in Suzuki Cross-Couplings in Aqueous Media. European

Journal of Organic Chemistry , (26), 5090-5099, S5090/1-S5090/38.

35) Martins, D. d. L.; Alvarez, H. M.; Aguiar, L. C. S., Microwave-assisted Suzuki reaction catalyzed by Pd(0)-PVP nanoparticles. Tetrahedron Letters 51, (52),

6814-6817.

36) Chattopadhyay, K.; Dey, R.; Ranu, B. C., Shape-dependent catalytic activity of copper oxide-supported Pd(0) nanoparticles for Suzuki and cyanation reactions. Tetrahedron Letters 2009, 50, (26), 3164-3167.

37) Astruc, D.; Ornelas, C.; Diallo, A. K.; Ruiz, J., Extremely efficient catalysis of carbon-carbon bond formation using "click" dendrimer-stabilized palladium nanoparticles. Molecules 15, 4947-4960.

38) Diallo, A. K.; Ornelas, C.; Salmon, L.; Aranzaes, J. R.; Astruc, D.,

"Homeopathic" catalytic activity and atom-leaching mechanism in Miyaura-

Suzuki reactions under ambient conditions with precise dendrimer-stabilized

Pd nanoparticles. Angewandte Chemie, International Edition 2007, 46, (45),

8644-8648.

39) Das, D. D.; Sayari, A., Applications of pore-expanded mesoporous silica 6.

Novel synthesis of monodispersed supported palladium nanoparticles and their catalytic activity for Suzuki reaction. Journal of Catalysis 2007, 246, (1), 60-

65.

89

40) Sullivan, J. A.; Flanagan, K. A.; Hain, H., Suzuki coupling activity of an aqueous phase Pd nanoparticle dispersion and a carbon nanotube/Pd nanoparticle composite. Catalysis Today 2009, 145, (1-2), 108-113.

41) Zhang, M.; Zhang, W., Pd Nanoparticles Immobilized on pH-Responsive and

Chelating Nanospheres as an Efficient and Recyclable Catalyst for Suzuki

Reaction in Water. Journal of Physical Chemistry C 2008, 112, (16), 6245-

6252.

42) Gude, K.; Narayanan, R., Synthesis and Characterization of Colloidal-

Supported Metal Nanoparticles as Potential Intermediate Nanocatalysts.

Journal of Physical Chemistry C 2010, 114, (14), 6356-6362.

43) Flaschsbart, H.; Stoeber, W., J. Coll. Interf. Sci. 1969, 30, 568-573.

44) Narayanan, R.; El-Sayed, M. A., Effect of Catalysis on the Stability of Metallic

Nanoparticles: Suzuki Reaction Catalyzed by PVP-Palladium Nanoparticles.

Journal of the American Chemical Society 2003, 125, (27), 8340-8347.

45) Li, Y.; Hong, X. M.; Collard, D. M.; El-Sayed, M. A., Org. Lett. 2000, 2,

2385.

46) Li, Y.; Boone, E.; El-Sayed, M. A., Langmuir 2002, 18, 4921.

47) Teranishi, T.; Miyake, M., Chem. Mater. 1998, 10, 594.

48) Narayanan, R.; El-Sayed, M. A., Effect of Colloidal Catalysis on the

Nanoparticle Size Distribution: Dendrimer-Pd vs PVP-Pd Nanoparticles

Catalyzing the Suzuki Coupling Reaction. Journal of Physical Chemistry B

2004, 108, (25), 8572-8580.

49) Narayanan, R.; El-Sayed, M. A., Effect of Colloidal Nanocatalysis on the

Metallic Nanoparticle Shape: The Suzuki Reaction. Langmuir 2005, 21, (5),

2027-2033.

50) Mills, G. A.; Boedeker, E. R.; Oblad, A. G., Chemical Characterization of

Catalysts. I. Poisoning of Cracking Catalysts by Nitrogen Compounds and

Potassium Ions. J. Am. Chem. Soc. 1950, 72, (4), 1554-1560.

90

51) Du, Y.; Chen, H.; Chen, R.; Xu, N., Poisoning effect of some nitrogen compounds on nano-sized nickel catalysts in p-nitrophenol hydrogenation.

Chm. Engin. J. 2006, 125, (1), 9-14.

52) Del Gallo, P.; PHam-Hou, C.; York, A. P. E.; Ledoux, M. J., Comparison of the Effects of Nitrogen Poisoning on Molybdenum Oxycarbide and Pt/Zeolite

Catalysts in the Isomerization of n-Heptane. Ind. Eng. Chem. Res. 1996, 35,

(10), 3302-3310.

53) Jana, S.; Haldar, S.; Koner, S., Heterogeneous Suzuki and Stille coupling reactions using highly efficient palladium(0) immobilized MCM-41 catalyst

Tetrahedron Letters 2009, 50, (34), 4820-4823.

54) Astruc, D., Palladium Nanoparticles as Efficient Green Homogeneous and

Heterogeneous Carbon # Carbon Coupling Precatalysts: A Unifying View.

Inorg. Chem. 2007, 46, (6), 1884-1894.

91

MANUSCRIPIT- III

Submitted for publication in Journal of Physical Chemistry C

Functionalized Colloidal Supported Metal Nanoparticles (CSMNs) as Efficient

Recyclable Nanocatalysts for Liquid-Phase Heck Reaction

Kalyani Gude and Radha Narayanan *

Department of Chemistry, University of Rhode Island, Kingston, RI, USA, 02881

92

Chapter 4

Functionalized Colloidal Supported Metal Nanoparticles (CSMNs) as Efficient

Recyclable Nanocatalysts for Liquid-Phase Heck Reaction

Abstract

Transition metal nanoparticles show excellent catalytic activity because of their small size. We investigate the catalytic activity, stability and recycling potential of two different types of nanocatalysts. The two catalysts used are the palladium nanoparticles that are supported onto the surface of functionalized and unfunctionalized silica colloids. We used aminopropyltriethoxysilane (APTES) as the linker to functionalize silica colloids. Both the catalysts are used to catalyze Heckreaction between styrene and iodobenzene that forms trans -stilbene. The fresh and recycled catalysts are characterized using TEM to observe any changes that might be taking place to the catalysts after its use in the reaction. Our studies showed that the palladium nanoparticles supported onto unfunctionalized silica colloid supports showed more catalytic activity than the palladium nanoparticles supported onto functionalized silica colloid supports during the first cycle of reaction. Our studies also showed that the palladium nanoparticles supported onto functionalized silica colloid supports showed more stability and recycling potential compared to palladium nanoparticles supported on unfunctionalized silica colloid supports.

Abbreviations

Transmission Electron Microscopy (TEM), (3-aminopropyl)triethoxysilane (APTES),

Polyvinylpyrrolidine (PVP), Poly(N-vinylimidazole) (PVI), Colloidal Supported

93

Metal Nanoparticles (CSMNs), High Performance Liquid Chromatography (HPLC)

(3-mercaptopropyl)trimethoxysilane (MPTMS)

Introduction

Transition metal nanoparticles are attractive catalysts as they have large surfaceto-volume ratio. The application of metal nanoparticles in catalysis is an important topic researchers have been working on in recent years because of their characteristic high surface to volume ratio. A greater number of surface atoms are exposed to reactant molecules that make them promising catalysts for some important industrial chemical reactions.

1-4 Traditionally available catalysts can be classified as homogeneous and heterogeneous catalysts. The homogeneous nanocatalysts include colloidal suspensions of transition metal nanoparticles 5-12 and the heterogeneous nanocatalysts include the colloidal suspensions of metal nanoparticles supported on supports such as polymers 13,14 metal oxides, 15,16 mesoporous silica, 17,18 carbon materials 19 and zeolites.

20 The direct use of metal nanoparticles (homogeneous) as catalysts is often difficult because of their high tendency to aggregate. Hence more research has been in progress to design a different kind of catalyst where the nanoparticles can be supported on solid matrix. The activity and selectivity of these catalysts depend on the nature and chemical structure of the support as well as the morphology of the supports. The presence of promoters and poisons and the interaction between the support and metal nanoparticles may also have an effect on catalysis by metal nanoparticles.

21,22

The palladium catalyzed reaction between an aryl halide and alkene was first

94

reported by Mizoroki 23 and Heck 24 in early 1970s. The Mizoroki-Heck reaction has been most widely used for the carbon-carbon bond formation reactions that allow alkylation, arylation, and vinylation of various alkenes through their reaction with alkyl, aryl and vinyl halides in the presence of palladium and base. This reaction is mostly carried out in the presence of palladium catalysts involving phosphine ligands.

But, these ligands are air-sensitive, poisonous, and unrecoverable from the reaction medium. Later, researchers tried to test phosphine-free catalytic systems and were successful.

25,26

Beller and Reetz for the first time showed that the Heck and Suzuki reactions can also be catalyzed by preformed palladium nanoparticles that are stabilized by tetraalkyl ammonium halides or PVP 27,28 . The problems with these homogeneous catalysts is extreme difficulty of separation, low recycling potential, loss of metal, or deactivation via aggregation of palladium nanoparticles formed in situ during Heck reaction. To overcome these problems, many chemists have tried designing heterogeneous catalytic systems. This system allows easy catalyst recovery, product separation, and inhibits loss of metal. In this regard, the homogeneous palladium nanoparticles are supported or immobilised on PVP-grafted silica, 29 PVI grafted silica 30 and amine functionalized mesoporous silica 31 and are used as catalysts for

Heck reaction. All the studies have shown leaching of palladium metal into the solution from the support.

The recycling potential is also an important factor to be considered while synthesizing the nanocatalysts because one wants to use it several times before losing its stability and catalytic activity. The recycling potential of both homogeneous and

95

heterogeneous catalysts is tested using Suzuki, Heck and many other kinds of reactions.

32,33 The studies using colloidal palladium nanoparticles showed poor recycling potential for the Suzuki reaction and the studies using palladium nanoparticles supported on activated carbon showed better recycling potential for upto five consecutive cycles.

34,35

Here, we discuss the colloidal silica supported metal nanoparticles that are used as catalysts for Heck reaction. As far as we know, there is no report in the literature where the PVP-Pd nanoparticles is supported on silica colloids and used in crosscoupling reactions. From our previous work, the colloidal supported palladium nanoparticles showed better catalytic activity for Suzuki reaction between phenylboronic acid and iodobenzene compared to the bulk silica supported metal nanoparticles 36 . The colloidal supported metal nanoparticles act as an intermediate kind of catalyst between homogeneous and heterogeneous catalytic systems. In this paper, we focused on testing the stability and recycling potential of colloidal supported metal nanoparticles.

The CSMNs have a high surface area and highly active surface atoms of the nanoparticles. They have the ability to be suspended in solution during catalysis and can be easily separated from the reaction mixture. They act as heterogeneous catalysts in a near homogeneous format. The palladium metal nanoparticles that are supported on functionalized silica colloids showed better stability and recycling potential compared to the palladium metal nanoparticles supported on unfunctionalized silica colloids. In the present study, APTES is chosen as a linker since it has a nitrogen group that the palladium nanoparticles can bind to and form the Pd-N bond. In the

96

case of unfunctionalized silica colloidal supports, palladium nanoparticles are adsorbed onto the surface of the support. In the case of APTES functionalized silica colloidal supports, palladium nanoparticles are covalently attached to the support.

During the first cycle of Heck reaction, unfunctionalized colloidal supported metal nanoparticles showed higher catalytic activity compared to functionalized colloidal supported metal nanoparticles. The catalytic activity decreased drastically during the second and third cycles of Heck reaction for unfunctionalized CSMNs (Colloidal

Supported Metal Nanoparticles). There is not much change in the catalytic activity exhibited by functionalized CSMNs during the second and third cycles of Heck reaction. The functionalised and unfunctionalised CSMNs before and after the reaction are characterized using TEM. TEM studies have shown that unfunctionalized CSMNs showed rapid aggregation of nanoparticles after their use for Heck reaction. The functionalized CSMNs showed an increase in the size of nanoparticles after their use in heck reaction and this increase can be attributed to the Ostwald ripening effect 37 .

Experimental Procedure

Synthesis of Palladium Nanoparticles

The palladium nanoparticles were synthesized using ethanol reduction method.

38,39 The palladium precursor solution is prepared by adding 0.0887g of PdCl

2 and 6ml of 0.2M HCl and then diluting to 250 mL of doubly deionized water. This mixture is sonicated for about an hour. A solution of 15ml of 2mM H

2

PdCl

4

, 21 mL of doubly deionized water, 0.0667g of PVP and 4 drops of 1M HCl was heated to 100

°

C.

97

When the solution began to reflux, 14 mL of ethanol was added and refluxed for 3 hours. This resulted in a brown colored suspension of palladium nanoparticles.

Synthesis of Silica Colloids

The silica colloids were synthesized using Stoeber synthesis method.

40 At first,

30ml of ethanol and 2.4 mL of ammonium hydroxide were added to an Erlenmeyer flask and the solution was stirred for 5 min. Then, 1.2 mL of tetraethylorthosilicate

(TEOS) was added to the solution containing ethanol and ammonium hydroxide and this solution was stirred overnight. The solution after the reaction consists of a suspension of silica colloids.

Functionalization of Silica Colloids

We used APTES as a linker to functionalize the silica colloids. To the suspension of silica colloids, 100 uL of 3-aminopropyltriethoxysilane (APTES) linker was added and reacted for 6hrs. We have also prepared unfunctionalized silica colloids for direct adsorption of palladium nanoparticles onto the surface of silica colloids.

Synthesis of Colloidal Supported Palladium Nanoparticles using Functionalized and

Unfunctionalised Silica Colloids

2 ml of functionalized and unfunctionalized silica colloids were added to 6ml of palladium nanoparticles in a scintillation vial and the suspension was stirred for 24 hrs. After 24 hrs, CSMNs, both functionalized and unfunctionalized, were centrifuged four times at 12,000 rpm for 3 minutes each time. During the first two centrifugation

98

cycles, they were redispersed in ethanol and during the last two centrifugation cycles, they were redispersed in doubly deionized water. After the final cycle of centrifugation, the supernatant was removed and the colloidal silica supported palladium nanoparticles were allowed to air-dry. 0.1g of the functionalized and unfunctionalized CSMNs are used as catalysts for Heck reaction.

Heck Reaction

The Heck reaction between iodobenzene and styrene was catalyzed using the synthesized CSMNs. A mixture of iodobenzene (1.0 mmol), styrene (2.0 mmol), triethylamine (1.5 mmol), DMF (20 mL) and 0.1g of catalyst was stirred at 120

°

C under inert atmosphere. The reaction is allowed to proceed for 2 hours. To study the progress of the reaction, the aliquots of samples were collected at different time intervals and quantified by HPLC analysis.

Recycling of CSMNs for the Second and Third cycles of Heck Reaction

After the first cycle of reaction, the CSMNs, both functionalized and unfunctionalized, were centrifuged six times at 12,000 rpm for 3 minutes each time.

During the first two cycles, they were redispersed in acetonitrile and during the next two cycles they were redispersed in ethanol and during the last two cycles they were redispersed in water. This whole process is done to remove any reactants and products that might be present on the surface of the catalyst. After the final cycle of centrifugation, the supernatant was removed and the CSMNs were allowed to air dry.

The same procedure is followed to recycle the catalyst after the second cycle of

99

reaction.

Characterization Studies with TEM to assess the Stability of Catalysts

In order to observe the changes in the nanoparticles and also the support, the catalyst samples before and after their use in the reaction are spotted on Formvarstabilized copper TEM grids. One drop of dilute suspensions of catalysts and recycled catalysts were used for spotting. The drop was allowed to air-dry for nearly 2 hours.

JEOL 2100EX TEM was used to obtain the images of the spotted samples. The

UTHSCSA ImageTool for Windows Version 3 image analysis software was used to determine the size distributions of the palladium nanoparticles supported on silica colloids.

Catalytic Activity Measurements using HPLC

HPLC measurements were conducted on Agilent 1120 Compact LC. A calibration curve for determining the concentration of trans -stilbene was constructed by plotting peak area versus concentration of trans -stilbene standards. The standards prepared were 1, 2.5, 5, 10, and 20 mM transstilbene in dimethylformamide. For

HPLC measurements, all the samples were diluted to 1/4 th or 1/8 th or 1/12 th of the original concentration, so that all the peak areas are within the range of the calibration curve. The actual concentration was determined by multiplying the concentration of the diluted sample with the dilution factor. The concentration of transstilbene formed during the first, second, and third cycles of the Heck reaction was investigated using

HPLC.

100

Results and Discussion

The CSMNs have a high surface area and highly active surface atoms of the nanoparticles. They have the ability to be suspended in solution during catalysis and can be easily separated from the reaction mixture. The CSMNs act as heterogeneous catalysts in a near homogeneous format. A careful four-step procedure is applied to synthesize colloidal supported metal nanoparticles. These steps include, synthesis of palladium nanoparticles, synthesis of silica colloids, functionalization of silica colloids and supporting the palladium nanoparticles onto functionalized and unfunctionalized silica colloids. In our previous study, we have used mercaptopropyltrimethoxysilane

(MPTMS) and aminopropyltriethoxysilane (APTES) as linkers to functionalize silica colloids. We tested the palladium nanoparticles supported on MPTMS and APTES functionalized silica colloids as catalysts for Suzuki reaction 41 . We found the palladium nanoparticles supported on MPTMS functionalized silica colloids showed no catalytic activity even after 24 hrs of running the reaction. But, the palladium nanoparticles supported on APTES functionalized silica colloids started showing the catalytic activity during the first hour of the reaction. Based on these results, we have chosen APTES as a linker to functionalize silica colloids and use them as catalysts for

Heck reaction.

We discuss the synthesis and characterization of two different types of colloidal supported metal nanoparticles (CSMNs). We also discuss the separation of the catalyst from the reaction mixture after first, second and third cycles of Heck-reaction and their characterization using TEM. We also discuss the results of the catalytic activity obtained by the fresh and recycled catalysts. HPLC is used to monitor the kinetics of

101

the Heck reaction between styrene and iodobenzene that forms trans -stilbene.

Synthesis and Characterization of Unfunctionalized CSMNs

The PVP capped palladium nanoparticles having an average size of 4.3 + 1.1 nm are adsorbed onto the surface of unfunctionalized silica colloids. Figure 1a shows the TEM image of PVP capped palladium nanoparticles and Figure 1b shows the size distribution histogram of PVP capped palladium nanoparticles obtained using the ethanol reduction method. It can be seen that the palladium nanoparticles have an average size of 4.3 + 1.1 nm. The palladium nanoparticles cannot be adsorbed on the inside of silica colloids, as the silica colloids are not porous; instead they are adsorbed on the surface of silica spheres. The palladium nanoparticles supported on unfunctionalized silica colloids are characterized using TEM. Figure 2a shows the

TEM image of palladium nanoparticles supported on the unfunctionalized silica colloids. Figure 2b shows the size distributions of the palladium nanoparticles supported on unfunctionalized silica colloids. The black specks in the TEM images are the palladium nanoparticles supported onto unfunctionalized silica colloids. It can be seen that the palladium nanoparticles supported onto unfunctionalized silica colloids have an average size of 4.3 + 0.9 nm.

Recycled Catalyst Characterization

The unfunctionalized CSMNs are recycled after the first, second and third cycles of Heck reaction and are characterized using TEM. Figure 3a shows the TEM image of palladium nanoparticles supported on unfunctionalized silica colloids after the first

102

cycle of reaction. Figure 3b shows the size distribution of the palladium nanoparticles supported on unfunctionalized silica colloids after the first cycle of reaction. It can be seen that the average size of the palladium nanoparticles increased to 10.5 + 2.9 nm.

By comparing the average size of the palladium nanoparticles and the standard deviation of palladium nanoparticles before and after the first cycle in Figure 2b and

Figure 3b, we can see that the size of the nanoparticles increased after the first cycle and also the standard deviation is increased too. The increase in the size of nanoparticles and also the increase in the standard deviation of the nanoparticles might be due to the Ostwald ripening process occurring under the reaction conditions.

Ostwald ripening is a process of growth of metal clusters. This process involves the detachment of metal atoms from the less stable metal clusters and reattachment onto the most stable metal clusters that causes the growth of the nanoparticles. Figure 4a shows the TEM image of the palladium nanoparticles supported on unfunctionalized silica colloids after the second cycle of Heck reaction. Figure 4b shows the TEM image of palladium nanoparticles supported on unfunctionalized silica colloids after the third cycle of Heck reaction. It can be seen that in both Figure 4a and Figure 4b, the palladium nanoparticles aggregated and precipitated which shows the instability of the catalyst. These two observations make sense because the catalytic activity decreased drastically after the first cycle of Heck-reaction that will be explained later.

Synthesis and Characterization of Functionalised CSMNs

The palladium nanoparticles are covalently attached to the surface of functionalized silica colloids. The surface of the silica colloids is functionalized using

103

100 uL of aminopropyl triethoxysilane (APTES). Figure 5a shows the TEM image of palladium nanoparticles supported onto APTES functionalized silica colloids. Figure

5b shows the size distributions of the palladium nanoparticles supported on APTES functionalised silica colloids. It can be seen that the palladium nanoparticles attached onto functionalized silica colloids have an average size of 4.1 + 0.8 nm.

Recycled Catalyst Characterisation

The functionalized CSMNs are recycled after the first, second and third cycles of

Heck reaction and characterized using TEM. Figure 6a shows the TEM image of palladium nanoparticles supported on APTES functionalized silica colloids after the first cycle of reaction. Figure 6b shows the size distributions of the palladium nanoparticles supported on functionalized silica colloids after the first cycle of Heck reaction. It can be seen that the average size of the palladium nanoparticles increased slightly to 4.8 nm with a standard deviation of 1.2 nm. By comparing the average size of the palladium nanoparticles and the standard deviation of palladium nanoparticles before and after the first cycle in Figure 5b and Figure 6b, we can see that the size of the nanoparticles increased slightly after the first cycle and also the standard deviation is increased slightly too. By comparing the size distributions and standard deviations of palladium nanoparticles supported on unfunctionalized silica colloids (Figure 3b) and functionalized silica colloids (Figure 6b) after the first cycle of Heck-reaction, we can see that there is a drastic increase in the size of nanoparticles that are adsorbed onto unfunctionalized silica colloids. This can be due to large amount of detachment of metal atoms from the smaller metal clusters attached onto unfunctionalized silica

104

colloids and reattachment onto the larger metal clusters. This suggests that the palladium nanoparticles are weakly adsorbed on the unfunctionalized silica colloids and are less stable. This also explains why the nanoparticles attached onto functionalized silica colloids are more stable compared to the ones adsorbed on unfunctionalized silica colloids. Figure 7a shows the TEM image of palladium nanoparticles supported on APTES functionalized silica colloids after the second cycle of reaction. Figure 7b shows the size distributions of the palladium nanoparticles supported on functionalized silica colloids after the second cycle of Heck reaction. It can be seen that the average size of the palladium nanoparticles increased to 6.3 nm with a standard deviation of 2.2 nm. By comparing the average size of the palladium nanoparticles and the standard deviation of palladium nanoparticles after the first and second cycle of the heck reaction, in Figure 6b and Figure 7b, we can see an increase in the size and standard deviation. This could be attributed to the Ostwald ripening process as described earlier. Figure 8a shows TEM image of palladium nanoparticles supported on APTES functionalized silica colloids after the third cycle of Heck reaction. Figure 8b shows the size distribution of the palladium nanoparticles supported on functionalized silica colloids after the third cycle of Heck reaction. It can be seen that the average size of the palladium nanoparticles increased to 6.8 nm with a standard deviation of 2.5 nm. This could also be attributed to Ostwald ripening process described earlier. Figure 7b and 8b shows that there are smaller nanoparticles as well as larger nanoparticles supported on the functionalized silica colloids and hence wide distribution.

105

Catalytic Activity Studies

We have tested the use of two different catalysts for the Heck reaction between iodobenzene and styrene that forms trans -stilbene. The catalysts used were functionalized and unfunctionalized CSMNs. The palladium content was same on both the catalysts since we prepared them in the same way. We have used reversed phase

HPLC to follow the kinetics of the Heck reaction for two hours. We will compare the catalytic activity of these two different catalysts during the first cycle of reaction. We will also compare the catalytic activity obtained by each individual catalyst for up to two reuses.

We have compared the catalytic activity of the palladium nanoparticles supported onto unfunctionalized silica colloids with that of palladium nanoparticles supported onto functionalized silica colloids. Figure 9 compares the catalytic activity of these two types of nanocatalysts based on the HPLC studies during the first cycle of the reaction. The catalytic activity of the palladium nanoparticles supported onto unfunctionalized silica colloids is a little higher than the palladium nanoparticles supported on functionalized silica colloids. In our previous report, where we used the same catalysts for Suzuki reaction between iodobenzene and phenylboronic acid, the same trend is observed. We reported that the lesser catalytic activity observed by the palladium nanoparticles that are supported onto APTES functionalized silica colloids could be due to the poisoning by the APTES linker. This could also be due to lower amount of detachment of metal atoms from the metal clusters that are covalently attached onto silica colloids. This also suggests that most of the catalytic activity exhibited by the functionalized CSMNs must be occurring heterogeneously. The

106

increased catalytic activity by the nanoparticles supported onto unfunctionalized silica colloids might be due to more detachment of metal atoms from the metal clusters that could participate in the catalytic cycle and attribute to the increase in catalytic activity.

Figure 10 compares the catalytic activity obtained by the fresh and recycled catalysts for up to three cycles of the reaction done with palladium nanoparticles supported onto unfunctionalized silica colloids. The first cycle of the reaction is done with the fresh catalyst, the second and third cycles of the reaction is done with reused catalysts.

There is a drastic decrease in the catalytic activity after the first cycle of the reaction.

The decrease in catalytic activity after the first cycle can be attributed to the increase in the average size of the nanoparticles after the first cycle of reaction that can be seen in figures 3a and 3b. Also there is a decrease in the catalytic activity from 2 nd cycle to

3 rd cycle. This can be explained based on the fact that the nanoparticles aggregated after the second cycle of the reaction that can be seen from figure 4a. Figure 11 compares the catalytic activity obtained by the fresh and recycled catalysts for up to three cycles of the reaction done with palladium nanoparticles supported onto APTES functionalized silica colloids. The first cycle of the reaction is done with the fresh catalyst, the second and third cycles of the reaction is done with reused catalysts.

There is not much decrease in the yield of transstilbene after the first and second cycles of the reaction that can be attributed to the stability of the nanoparticles covalently attached onto APTES functionalized silica colloids. This also can be explained based on the fact that there is not much increase in the size of the nanoparticles after the first cycle of the reaction. This suggests that the palladium

107

nanoparticles covalently attached onto functionalized silica colloids are more stable compared to the ones adsorbed on unfunctionalized silica colloids.

Conclusions

The palladium nanoparticles supported on unfunctionalized silica colloids showed a greater increase in size compared to the ones supported on functionalized silica colloids after the first cycle of Heck reaction. There is greater aggregate formation of palladium nanoparticles that are supported on unfunctionalized silica colloids after the second and third cycles of Heck-reaction. Also, the palladium nanoparticles supported on unfunctionalized silica colloids showed a decrease in catalytic activity during the second and third cycles of Heck reaction. The palladium nanoparticles supported on functionalized silica colloids showed an increase in size after the first, second and third cycles of Heck-reaction, but there is no aggregation of nanoparticles even after the third cycle of Heck-reaction. There is not much decrease in the catalytic activity exhibited by the functionalized CSMNs after the first and second cycles of Heck-reaction. The catalytic activity exhibited by the palladium nanoparticles supported onto unfunctionalized silica colloids is higher than the catalytic activity exhibited by the ones supported on APTES functionalized silica colloids during the first cycle of Heck reaction. But, the stability and recycling potential of the palladium nanoparticles supported on functionalized silica colloids is better than the ones supported on unfunctionalized silica colloids. Based on the results obtained, it can be seen that palladium nanoparticles supported on functionalized

108

colloidal silica supports are better catalysts for Heck reaction between iodobenzene and styrene compared to the ones supported on unfunctionalized colloidal silica.

109

1b

4.3+1.1nm

40

30

20

10

0

0 2 4 6 8 10

!

Size of Nanoparticles (nm)

Figure 4-1. TEM image of the palladium nanoparticles (1a) Size distribution of palladium nanoparticles (1b)

!" n2 2b 4.3 ± 0.9nm

!

nm200 nm

Figure 4-2.

TEM image showing the palladium nanoparticles supported onto unfunctionalized silica colloid suspensions (a) Size distribution of palladium nanoparticles supported on unfunctionalised silica colloids (b)

110

3b

10.5 ± 2.9nm

!

Figure 4-3.

TEM image showing the palladium nanoparticles supported onto unfunctionalized silica colloid suspensions after the first cycle of heck reaction (a)

Size distribution of palladium nanoparticles supported on unfunctionalised silica colloids after the first cycle of heck reaction (b)

!

Figure 4-4.

TEM image showing the palladium nanoparticles supported onto unfunctionalized silica colloid suspensions after the second cycle of heck reaction (a)

TEM image showing the palladium nanoparticles supported onto unfunctionalized silica colloid suspensions after the third cycle of heck reaction (b)

111

5a

5b

4.1 ± 0.8 nm

80

60

40

20

200

0

0 1 2 3 4 5 6 7

Size of nanoparticles (nm)

8 9 10

Figure 4-5.

TEM image showing the palladium nanoparticles supported onto APTES functionalized silica colloid suspensions (a) Size distribution of palladium nanoparticles supported onto APTES functionalized silica colloids (b)

6b

4.8 ± 1.2 nm

35

30

25

20

15

10

5

0

0 1 2 3 4 5 6 7

Size of Nanoparticles (nm)

8 9 10

!

Figure 4-6.

TEM image showing the palladium nanoparticles supported onto APTES functionalized silica colloid suspensions after the first cycle of heck reaction (a) Size distribution of palladium nanoparticles supported onto APTES functionalized silica colloids after the first cycle of heck reaction (b)

112

7

6

5

4

7b

6.3 ± 2.2 nm

3

2

1

200 nm

0

0 2 4 6 8 10 12 14 16 18 20

Size of the Nanoparticles (nm)

!

!

Figure 4-7.

TEM image showing the palladium nanoparticles supported onto APTES functionalized silica colloid suspensions after the second cycle of heck reaction (a)

Size distribution of palladium nanoparticles supported onto APTES functionalized silica colloids after the second cycle of heck reaction (b)

8b 6.8 + 2.5 nm

8

6

4

2

0

0 2 4 6 8 10 12 14 16 18 20

Size of nanoparticles (nm)

!

!

Figure 4-8.

TEM image showing the palladium nanoparticles supported onto APTES functionalized silica colloid suspensions after the third cycle of heck reaction (a) Size distribution of palladium nanoparticles supported onto APTES functionalized silica colloids after the third cycle of heck reaction (b)

113

0.45

0.40

0.35

0.30

0.25

0.20

0.15

0.10

0.05

CSMNs No linker

CSMNs Amine linker

0.00

0 20 40 60 80 100 120

Time (min)

Figure 4-9. Comparisons of the catalytic activity obtained with functionalized and unfunctionalized CSMNs

0.45

0.40

0.35

0.30

0.25

0.20

0.15

0.10

0.05

0.00

0

1st Cycle

2nd Cycle

3rd Cycle

20 40 60 80 100 120

Time (min)

Figure 4-10. Comparisons of the catalytic activity with fresh and reused catalysts of palladium nanoparticles supported onto unfunctionalized silica colloids

114

0.45

0.40

0.35

0.30

0.25

0.20

0.15

0.10

0.05

0.00

0

1st Cycle

2nd Cycle

3rd Cycle

20 40 60 80 100 120

Time ( min)

Figure 4-11. Comparisons of the catalytic activity with fresh and reused catalysts of palladium nanoparticles supported onto functionalized silica colloids

115

References

1) Lewis, L. N. Chem. Rev . 1993 , 93 , 2693-2730.

2) Roucoux, A.; Schulz, J.; Patin, H. Chem. Rev . 2002 , 102 , 3757-3778.

3) Min, B. K.; Friend, C. M. Chem Rev . 2007 , 107, 2709-2724.

4) Arcadi, A.; Chem Rev. 2008 , 108, 3266

5) Astruc, D. Tetrahedron: Asymmetry. 2010 , 21 , 1041 1054.

6) Boennemann, H.; Nagabhushana, K. S. Surf. Nanomol. Catal . 2006 , 63–94.

7) Boennemann, H.; Nagabhushana, K. S.; Richards, R. M. Nanopart. Catal.

2008 , 49–91.

8) Brock, S. L.; Senevirathne, K. Solid State Chem . 2008 , 181 (7), 1552–1559.

9) Bronstein, L. M.; Matveeva, V. G.; Sulman, E. M. Nanopart. Catal . 2008 , 93-

127

10) Duran Pachon, L.; Rothenberg, G. Appl. Organomet. Chem . 2008 , 22 (6), 288–

299

11) Gates, B. C. Nature Nanotechnol . 2008 , 3 (10), 583–584.

12) Gorodetskii, V. V.; Sametova, A. A.; Matveev, A. V.; Tapilin, V. M. Catal.

Today. 2009 , 144 (3-4), 219–234.

13) Kralik, M.; Biffis, A. J. Mol. Catal. A.

2001 , 177 , 113.

14) Akagi, T.; Baba, M.; Akashi, M. Polymer . 2007 , 48, 6729

15) Risse, T.; Shaikhutdinov, S.; Nilius, N.; Sterrer, M.; Freund, H. J. Acc. Chem.

Res . 2008 , 41 , 949

16) Chen, M.S.; Goodman, D. W. Chem. Soc. Rev . 2008 , 37 , 1860

17) Stein, A.; Adv. Mater , 2003 , 15 , 763

18) Bronstein, L.M.; Top. Curr. Chem.; 2003, 266, 55

19) Xiaoli, J.; Bo, H.; Jigen, M.; Junhua. Y.; Qixian, Z.; Li, N. Carbon , 2012 , 50 ,

3083-3091

116

20) Marcelo, D. E.; Consuelo, H. M.; Maria, N. T.; Yolanda, P. Catalysis Today,

2011 , 172 , 13-20

21) Wan, Y.; Wang, H.Y.; Zhao, Q.; Klingstedt, M.; Terasaki, O.; Zhao, D.

J.Am.Chem. Soc . 2009, 131, 4541-4550

22) Mehnert, C. P.; Weaver, D. W.; Ying, J.Y. J. Am. Chem. Soc.

1998 , 120,

12289

23) Mizoroki, T.; Mori, K.; Ozaki, A. Bull. Chem. Soc. Jpn . 1971 , 44 , 581

24) Heck, R. F.; Nolley, J. P . J. Org. Chem . 1972 , 37 , 2320

25) Alo, B. I.; Kandil, A.; Patil, P. A.; Sharp, M. J.; Siddiqui, M. A.; Snieckus, V.

J. Org. Chem.

1991 , 56 , 3763

26) Wallow, T. I.; Novak, B. M. J. Org. Chem.

1994 , 59 , 5034

27) Belleer, M.; Fischer. H.; Kuhlein, K.; Reisinger. C. P.; Herrmann, W. A.;

J.Organomet. Chem . 1996 , 520 , 257

28) Reetz, M.T.;Breinbauer, R.; Wanninger, K.; Tetrahedron Lett , 1996 , 37 , 4499

29) Tamami, B.; Allahyari, H.; Ghasemi, S.; Farjadian, F.; Journal of

Organometallic Chemistr y. 2011 , 696, 594-599

30) Tamami, B.; Allahyari, H.; Ghasemi, S.; Farjadian, F.; Sohelia, G.; Iranian

Polymer Journal, 2011, 20, 699-712

31) Peiyu, W.; Qingshan, L.; Jiangong, L. Materials Research Bulletin.

2010 , 45 , 129-134

32) Feng, Y.; Li, L.; Li, Y.; Zhao, W.; Gu, J. Shi, J. Journal. Of. Molecular.

Catalysis.

2010 , 322, 50-54

33) Calo, V.; Nacci, A.; Monopoli, A.; Montingelli, F.; J Org Chem , 2006 , 71 ,

4339

34) Narayanan, R.; El-Sayed, M. A.; J Am Chem Soc 2003, 12 (27), 8340-8347

35) Narayanan, R.; El-Sayed, M. A.; J Catal , 2005 , 234(2) , 348

36) Gude, K.; Narayanan, R.; J. Phys. Chem C . 2011 , 115 , 12716-12725

37) Howard, A.; Mitchell, C. E. J.; Edgell, R. G. Surf. Science . 2002 , 515 , L504

117

38) Li, Y.; Hong, X. M.; Collard, D. M.; El-Sayed, M. A., Org. Lett. 2000, 2, 2385

39) Li, Y.; Boone, E.; El-Sayed, M. A., Langmuir 2002, 18, 4921

40) Flaschsbart, H.; Stoeber, W., J. Coll. Interf. Sci. 1969, 30, 568-573

41) Gude, K.; Narayanan, R.; J. Phys. Chem C . 2010 , 114 , 6356-6362

118

BIBILOGRAPHY

Ahmadi, T. S.; Wang, Z. L.; Green, T. C.; Henglein, A.; El-Sayed, M. A. Science ,

1996 , 272 , 1924.

Aiken III, J. D.; Finke, R. G.; J. Mol. Catal. A: Chem. 1999 , 145 , 1-44.

Akagi, T.; Baba, M.; Akashi, M. Polymer . 2007 , 48, 6729.

Albers, P. W.; Konrad, M.; Frost, C. D.; Parker, S. F. Journal of Physical

Chemistry C . 2011 , 115(50), 24485-24493

Alo, B. I.; Kandil, A.; Patil, P. A.; Sharp, M. J.; Siddiqui, M. A.; Snieckus, V. J.

Org. Chem.

1991 , 56 , 3763

Amiens, C.; Decaro, D.; Chaudret, B.; Bradley, J. S.; Mazel, R.; Roucau, C. J. Am.

Chem. Soc. 1993 , 115 , 11638-11639

Antolini, E.; Salgado, J. R. C.; da Silva, R. M.; Gonzalez, E. R. Mat. Chem. Phys.

2007 , 101 , 395-403.

Arcadi, A.; Chem Rev. 2008 , 108, 3266

Arenz, M.; Landman, U.; Heiz, U. ChemPhysChem 2006, 7, (9), 1871-1879.

Asanithi, P.; Chaiyakun, S.; Limsuwan, P.; Journal of Nanomaterials . 2012 , 2012 ,

963609.

Astruc, D. Tetrahedron: Asymmetry. 2010 , 21 , 1041 1054.

Astruc, D. Inorg. Chem. 2007, 46, (6), 1884-1894.

Astruc, D.; Lu, F.; Aranzaes, J. R. Ange. Chem., Int. Ed.

2005 , 44 , 7852-7872.

Astruc, D.; Ornelas, C.; Diallo, A. K.; Ruiz, J. Molecules 15, 4947-4960.

Badley, R. D.; Ford, W. T.; McEnroe, F. J.; Assink, R. A. Langmuir 1990 , 6 , 792

Baeumer, M.; Libuda, J.; Neyman, K. M.; Roesch, N.; Rupprechter, G.; Freund,

H.-J. Phys. Chem. Chem. Phys.

2007 , 9 , 3541-3558

Balogh, L.; Tomalia, D. A. J. Am. Chem. Soc. 1998 , 120 , 7355

Bard, A. J.; Faulkner, L.R. Electrochemical Methods: Fundamentals and

Applications , Wiley, New York, 1980

119

Barkhuizen, D.; Mabaso, I.; Viljoen, E.; Welker, C.; Claeys, M.; van Steen, E.;

Fletcher, J. C. Q. Pure and Applied Chemistry 2006, 78, (9), 1759-1769.

Baykal, A.; Toprak, M. S.; Durmus, Z.; Sozeri, M. S. H.; Demir, A. J. Supercond.

Nov. Magn.

2012 , 25 , 1541-1549

Beller, M.; Fischer. H.; Kuhlein, K.; Reisinger. C. P.; Herrmann, W. A.;

J.Organomet. Chem . 1996 , 520 , 257

Bitter, J. H. Nanocatalysis 2006 , 99-125.

Boennemann, H.; Nagabhushana, K. S. Surf. Nanomol. Catal . 2006 , 63–94.

Boennemann, H.; Nagabhushana, K. S.; Richards, R. M. Nanopart. Catal.

2008 ,

49–91.

Boisselier, E.; Diallo, A. K.; Salmon, L.; Ornelas, C.; Ruiz, J.; Astruc, D.; J. Am.

Chem. Soc.

2010 , 132 , 2729–2742

Bonet, F.; Delmas, V.; Grugeon, S.; Urbina, H. R.; Silvert, P.; Tekaia-Elhsissen,

K. Nanostruct. Mater . 1999 , 11 , 1277-1284

Borsla, A.; Wilhelm, A. M.; Delmas, H. Catal. Today . 2001 , 66(2-4) , 389

Bottini, M.; Dawson, M. I.; Mustelin, T., Carbon nanotube-based inorganic composites. Chemistry of Carbon Nanotubes 2008, 3, 25-34.

Bouchat, V.; Feron, O.; Gallez, B.; Masereel, B.; Michiels, C.; Borght, T. V.;

Lucas, S.; Surface & Coatings Technology.

2011 , 205 , S577-S581

Brock, S. L.; Senevirathne, K. Solid State Chem . 2008 , 181 (7), 1552–1559.

Bronstein, L. M.; Matveeva, V. G.; Sulman, E. M. Nanopart. Catal . 2008 , 93-127

Bronstein, L.M.; Top. Curr. Chem.; 2003 , 266, 55

Budarin, V.; Clark, J. H.; Luque, R.; White, R. Material Matters (Milwaukee, WI,

United States) 2009, 4, (1), 19-22.

Bunker, C. E.; Novak, K. C.; Guliants, E. A.; Harruff, B. A.; Meziani, M. J.; Lin,

Y.; Sun,Y. P. Langmuir. 2007 , 23 , 10342-10347

Burda, C.; Chen, X.; Narayanan, R.; El-Sayed, M. A. Chem. Rev. (Washington,

DC, United States) 2005 , 105 , 1025-1102.

Bynghoon, Y.; Horng-Bin, P.; Wai, C. M. Journal of Physical Chemistry C.

2009,

113(4), 1520-1525

120

Calo, V.; Nacci, A.; Monopoli, A.; Montingelli, F.; J Org Chem , 2006 , 71 , 4339

Campelo, J. M.; Conesa, T. D.; Gracia, M. J.; Jurado, M. J.; Luque, R.; Marinas, J.

M.; Romero, A. A. Green Chem. 2008 , 10 , 853-858

Campelo, J. M.; Lee, A. F.; Luque, R.; Luna, D.; Marinas, J. M.; Romero, A. A.

Chem. Eur. J. 2008 , 14 , 5988-5995

Campelo, J. M.; Luna, D.; Luque, R.; Marinas, J. M.; Romero, A. A. Chem Sus

Chem 2009, 2, (1), 18-45.

Chan, K.-Y. Small Fuel Cells (5th Edition) 2004 , 305-334.

Chan, K.-Y.; Chi, N.; Phillips, D. L. Curr. Top. Electrochem.

1998 , 6 , 189-197

Chan, K.-Y.; Ding, J.; Ren, J.; Cheng, S.; Tsang, K. Y. J. Mat. Chem.

2004 , 14 ,

505-516.

Chang, J.-S.; Hwang, J.-S.; Park, S.-E. Res. Chem. Interm.

2003 , 29 , 921-938.

Chattopadhyay, K.; Dey, R.; Ranu, B. C. Tetrahedron Letters 2009, 50, (26),

3164-3167.

Chechik, V.; Crooks, R. M. J. Am. Chem. Soc. 2000 , 122 , 1243

Chen, C. L.; Mou, C. Y. Nanotech. Catal.

2004 , 1 , 313-327.

Chen, C.; Chen, M.; Serizawa, T.; Akashi, M. Chem. Communications .

1998 , 7 , 831-832

Chen, C.; Serizawa, T.; Akashi, M. Chem. Mater. 1999 , 11 , 1381

Chen, M.; Goodman, D. W. Chem. Soc. Rev.

2008 , 37 , 1860-1870.

Chen, X.; Hou, Y.; Wang, H.; Cao, Y.; He, J. Journal of Physical Chemistry C

2008, 112, (22), 8172-8176.

Choi, H.; Al-Abed, S. R.; Agarwal, S.; Dionysiou, D. D. Chem. Mater. 2008 , 20 ,

3649-3655

Chung, B. X.; Liu, C. P.; Materials Letters . 2004 , 58 , 1437-1440

Corain, B.; Centomo, P.; Lora, S.; Kralik, M. J. Molec. Catal. A: Chem.

2003 ,

204-205 , 755-762.

121

Corma, A.; Garcia, H. Chemical Society Reviews 2008, 37, (9), 2096-2126.

Crooks, R. M.; Zhao, M. Q.; Sun, L.; Chechik, V.; Yeung, L. K. Acc. Chem. Res.

2001 , 34 ,181-190

Cuenya, B. R. Thin Solid Films 518, (12), 3127-3150.

Curri, M. L.; Comparelli, R.; Striccoli, M.; Agostiano, A. Physical Chemistry

Chemical Physics 12, (37), 11197-11207.

Das, D. D.; Sayari, A. Journal of Catalysis 2007, 246, (1), 60-65.

Dasog, M.; Scott, R. W. J. Langmuir. 2007 , 23 , 3381-3387

Del Gallo, P.; PHam-Hou, C.; York, A. P. E.; Ledoux, M. J. Ind. Eng. Chem. Res.

1996, 35, (10), 3302-3310.

Diallo, A. K.; Ornelas, C.; Salmon, L.; Aranzaes, J. R.; Astruc, D. Angewandte

Chemie, International Edition 2007, 46, (45), 8644-8648.

Doris, R.; Claudio, M.; Fierro, J. L. G.; Patricio, R. Journal of the Chilean

Chemical Society.

2012 , 57(4), 1394-1399

Du, Y.; Chen, H.; Chen, R.; Xu, N. Chm. Engin. J. 2006, 125, (1), 9-14.

Dummer, N. F.; Salem, B.; James, H.; Robert, J.; Hutchings, G. J. Catalysis

Today. 2011 , 160(1), 50-54

Duran Pachon, L.; Rothenberg, G. Appl. Organomet. Chem . 2008 , 22 (6), 288–299

Duteil, A.; Schmid, G.; Meyer-Zaika, W. Chem. Commun. 1995 , 31-32

Dutta, P.; Sarkara, A. Adv. Synth. Catal . 2011 , 353 , 2814 – 2822

Eppler, A.; Rupprechter, G.; Guczi, L.; Somorjai, G. A. J. Phys. Chem. B 1997 ,

101 , 9973-9977.

Ershov, B. G.; Abkhalimov, E. V.

Colloid Journal, 2007 , 69

Ershov, B. G.; Janata, E.; Michaelis, M.; Henglein, A. J. Phys. Chem. 1991 , 95 ,

8996-8999

Ertl, G.; Knozinger, H.; Weitkamp, J. Handbook of Heterogeneous Catalysis ;

Wiley/VCH: New York/Weinheim, 1997

Fan, J.; Gao, Y. Journal of Experimental Nanoscience 2006, 1, (1-4), 457-475.

122

Feng, Y.; Li, L.; Li, Y.; Zhao, W.; Gu, J. Shi, J. Journal. Of. Molecular. Catalysis.

2010 , 322, 50-54

Fernando, K. A. S.; Smith, M. J.; Harruff, B. A.; Lewis, W. K.; Guliants, E. A.;

Bunker, C. E. J. Phys. Chem. C. 2009 , 113 , 500-503

Flaschsbart, H.; Stoeber, W. J. Coll. Interf. Sci.

1969 , 30 , 568-573.

Fukuoka, A.; Dhepe, P. L. Chemical Record 2009, 9, (4), 224-235.

Furlong, D. N.; Launikonis, A.; Sasse, W. H. F.; Sanders, J. V. J. Chem. Soc.,

1984 , 80 , 571-588

Gaffet, E.; Tachikart, M.; Kedim, O. E.; Rahouadj, R. Mater.Charact.

1996 , 36 ,

185-190

Ganapati, Y. D. Catal. Sur v . Asia 2005 , 9 , 117–137

Gates, B. C. Nature Nanotechnol . 2008 , 3 (10), 583–584.

Gianmario, M.; Luca, B.; Claudio, G.; Raimondo, M.; Giovanni, S.; Salvatore, C.

Catalysis Letters. 2009 , 132(1-2), 50-57

Gonzalez, C. M.; Liu, Y.; Scaiano, J. C. J. Phys. Chem. C. 2009 , 113 , 11861-

11867

Gorodetskii, V. V.; Sametova, A. A.; Matveev, A. V.; Tapilin, V. M. Catalysis

Today. 2009, 144, (3-4), 219-234.

Gubin, S. P.; Koksharov, Y. A.; Khomutov, G. B.; Yurkov, G. Y. Russian chemical reviews . 2005 , 74 (6) 489-520

Gude, K.; Narayanan, R. Journal of Physical Chemistry C. 2010, 114, (14), 6356-

6362.

Gude, K.; Narayanan, R. Journal of Physical Chemistry C .

2011 , 115 , 12716-

12725

Guoping, L.; Wenting, L.; Yunjun, L.; Min, X.; Chunpeng, C.; Xiaoqing, W. Chin.

J. Chem. 2012 , 30 , 541-546

Guzmán, M. G.; Dille, J.; Godet, S. International Journal of Chemical and

Biological Engineering. 2009, 2:3 , 104-111

123

Haifeng, Y.; Qingyi, L.; Feng, G.; Qihui, S.; Yan, Y.; Fuqiang, Z.; Songhai, X.;

Bo, T.; Dongyuan, Z. Advanced Functional Materials. 2005 , 15(8), 1377-1384

Harriman, A.; Thomas, J. M.; Millward, G. R. New J. Chem. 1987 , 11 , 757-762

Heck, R. F.; Nolley, J. P . J. Org. Chem . 1972 , 37 , 2320

Hermans, S.; Khimyak, T.; Raja, R.; Sankar, G.; Johnson, B. F. G.; Thomas, J. M.

Stud. Surf. Sci. Catal.

2003 , 145 , 181-184.

Hirai, H.; Ohtaki, M.; Komiyama, M. Chem. Lett.

1986 , 3 , 269-72

Hirai, H.; Ohtaki, M.; Komiyama, M. Chem. Lett. 1987 , 1 , 149-51

Hirai, H.; Yakurat, N. Polym. Adv. Tech.

2001 , 12(11-12) , 724

Hirtzel, C. S.; Rajagopalan, R. Colloidal Phenomena: Advanced Topics , Noyes

Publications, Park Ridge, NJ, 1985

Hou, W. B.; Dasog, M.; Scott, R. W. J. Langmuir. 2009 , 25 , 12954-12961

Howard, A.; Mitchell, C. E. J.; Edgell, R. G. Surf. Science . 2002 , 515 , L504

Hunter, R. J. Foundations of Colloid Science, vol. 1, Oxford University Press,

New York, 1986

Hunter, R. J. In Foundations of Colloid Science ; Oxford University Press: New

York, 1987; Vol. 1, p 316

Jacinto, M. J.; Santos, O. H. C. F.; Jardim, R. F.; Landers, R.; Rossi, Liane, M.

Applied Catalysis, A: General.

2009 , 360(2), 177-182

Jana, S.; Haldar, S.; Koner, S.

Tetrahedron Letters 2009, 50, (34), 4820-4823.

Jong-San, C.; Jin-Soo, H.; Sang-Eon, P. Res. Chem. Intermed.

2003 , 29 , 921-938

Karousis, N.; Tsotsou, G.-E.; Evangelista, F.; Rudolf, P.; Ragoussis, N.;

Tagmatarchis, N. Journal of Physical Chemistry C 2008, 112, (35), 13463-13469.

Kasemo, B.; Johansson, S.; Persson, H.; Thormahlen, P.; Zhdanov, V. P. Topics in

Catalysis 2000, 13, (1,2), 43-53.

Khanna, P. K.; Journal of Synthesis and Reactivity in Inorganic, Metal-Organic, and Nano- Metal Chemistry . 2007 , 37, 805-808

Kopple, K.; Meyerstein, D.; Meisel, D. J. Phys. Chem. 1980 , 84 , 870

124

Kralik, M.; Biffis, A. J. Mol. Catal. A.

2001 , 177 , 113.

Kumar, R.; Ghosh, A.; Patra, C. R.; Mukherjee, P.; Sastry, M. Nanotech. Catal.

2004 , 1 , 111-136

Kwong-Yu, C.; Jie, D.; Jiawen, R.; Shaoan, C.; Ying, T. K. J. Mater.

Chem.

2004 ,

14 , 505-516

Kwong-Yu, C.; Small Fuel Cells (5th Ed.) 2004 , 305-334

Lakina, N. V.; Sul'man, E. M.; Matveeva, V. G.; Mikhailov, I. A, Pharm. Chem. J.

2000 , 34(3) , 138

Laura, R.; Christian, P.; Philippe, S.; Montserrat, G.; Emmanuelle, T.

ChemCatChem, 2011 , 3(4), 749-754

Lewis, L. N. Chem. Rev . 1993 , 93 , 2693-2730.

Li, Y.; Boone, E.; El-Sayed, M. A. Langmuir 2002 , 18 , 4921.

Li, Y.; El-Sayed, M. A. J. Phys. Chem. B 2001 , 105 , 8938-8943

Li, Y.; Hong, X. M.; Collard, D. M.; El-Sayed, M. A. Org. Lett. 2000 , 2(15) , 2385

Li, Y.; Liu, J.; Wang, Y.; Wang, Z. L.; Chem. Mater.

2001 , 13, 1008-1014

Li, Y.; Petroski, J.; El-Sayed, M. A. J. Phys. Chem. B , 2000 , 104(47) , 10956

Lin, Y.; Finke, R. G. J. Am. Chem. Soc. 1994 , 116 , 8335

Liu, Z.; Lee, J. Y.; Chen, W.; Han, M.; Gan, L. M. Langmuir. 2004 , 20(1), 181

Liu, Z.; Ling, X. Y.; Lee, J. Y.; Su, X.; Gan, L. M. J. Mat. Chem.

2003 , 13(12),

3049

Logvinenko, V.; Polunina, O.; Mikhailov, Y.; Mikhailov, K.; and Bokhonov, B.;

Journal of Thermal Analysis and Calorimetry.

2007 , 90, 813-816

Lu, F.; Ruiz, J.; Astruc, D. Tetrahedron Lett. 2004 , 45 , 9443-9445

Marcelo, D. E.; Consuelo, H. M.; Maria, N. T.; Yolanda, P. Catalysis

Today, 2011 , 172 , 13-20

Marignier, J. L.; Belloni, J.; Delcourt, M. O.; Chevalier, J. P. Nature 1985 , 317 ,

344-345

125

Marin, M. L.; McGilvray, K. L.; Scaiano, J. C. J. Am. Chem. Soc. 2008 , 130 ,

16572-16584

Martins, D. d. L.; Alvarez, H. M.; Aguiar, L. C. S. Tetrahedron Letters 51, (52),

6814-6817.

Masatake, H.; Susumu, T.; Tetsuhiko, K.; Hiroyuki, K.; Genet, M. J.; Bernard, D.;

Journal of Catalysis. 1993 , 144(1), 175-192

McGilvray, K. L.; Decan, M. R.; Wang, D. S.; Scaiano, J. C. J. Am. Chem. Soc.

2006 , 128 , 15980-15981

Mehnert, C. P.; Weaver, D. W.; Ying, J.Y. J. Am. Chem. Soc.

1998 , 120, 12289

Mehran, G.; Donya, V.; Zahra, S. Chinese Chemical Letters. 2012 , 23(8) , 887-890

Mejias, N.; Pleixats, R.; Shafir, A.; Medio-Simon, M.; Asensio, G. European

Journal of Organic Chemistry , (26), 5090-5099, S5090/1-S5090/38.

Michaelis, M.; Henglein, A. J. Phys. Chem. 1992 , 96 , 4719-4724

Mills, G. A.; Boedeker, E. R.; Oblad, A. G. J. Am. Chem. Soc. 1950, 72, (4), 1554-

1560.

Min, B. K.; Friend, C. M. Chem Rev . 2007 , 107, 2709-2724.

Mingshu, C.; Wayne, G. D. Chem. Soc. Re V .

2008 , 37 , 1860-1870

Mizoroki, T.; Mori, K.; Ozaki, A. Bull. Chem. Soc. Jpn . 1971 , 44 , 581

Miyaura, N.; Yamada, K.; Suzuki. A. Tetrahedron Lett.

1979 , 20 , 3437

Mizuki, T.; Rajaram, B.; Seitaro, N.; Yasuhiro, I. Applied Catalysis, A: General.

2006 , 307(1), 78-84

Molenbroek, A. M.; Helveg, S.; Topsoe, H.; Clausen, B. S. Topics in Catalysis

2009, 52, (10), 1303-1311.

Mosseri, S.; Henglein, A.; Janata, E. J. Phys. Chem. 1989 , 93 , 6791-6795

Mueller, C.; Nijkamp, M. G.; Vogt, D. Eur. J. Inorg. Chem.

2005 , 4011-4021.

Nakagawa, K.; Yamagishi, M.; Nishimoto, H.; Ikenaga, N.; Suzuki, T.; Kobayashi,

T.; Nishitani-Gamo, M.; Ando, T. Chem. Mat . 2003 , 15(24), 4571

126

Napper, D. H. In Polymeric Stabilization of Colloidal Dispersions ; Academic

Press: London, 1983

Narayanan, R. Molecules 15, 2124-2138.

Narayanan, R.; El-Sayed, M. A. Chim. Oggi 2007 , 25 , 84-86.

Narayanan, R.; El-Sayed, M. A. J. Amer. Chem. Soc.

2003 , 125 , 8340-8347.

Narayanan, R.; El-Sayed, M. A. J. Amer. Chem. Soc.

2004 , 126 , 7194-7195.

Narayanan, R.; El-Sayed, M. A. J. Catal.

2005 , 234 , 348-355.

Narayanan, R.; El-Sayed, M. A. J. Phys. Chem. B.

2004 , 108 , 5726-5733.

Narayanan, R.; El-Sayed, M. A. J. Phys. Chem. B . 2004 , 108 , 8572-8580.

Narayanan, R.; El-Sayed, M. A. J. Phys. Chem. B.

2005 , 109 , 12663-12676

Narayanan, R.; El-Sayed, M. A. J. Phys. Chem. B . 2003 , 107(45) , 12416

Narayanan, R.; El-Sayed, M. A. Langmuir.

2005 , 21 , 2027-2033

Narayanan, R.; El-Sayed, M. A. Nano Lett.

2004 , 4 , 1343-1348.

Narayanan, R.; El-Sayed, M. A. Top. Catal. 2008 , 47 , 15-21

Narayanan, R.; Tabor, C.; El-Sayed, M. A. Top. Catal.

2008 , 48 , 60-74.

Neelgund, G. M.; Aderemi, O. Applied Catalysis, A: General. 2011 , 399(1-

2), 154-160

Oemer, D.; Olga, S.; Neil, C.; Ozin, G. A.; Journal of Materials Chemistry . 2003,

13(2), 328-334

Oki, M.; Rebrov, E. V.; Angel, B.; de Croon, M. H. J. M.; Schouten, J. C. Applied

Catalysis, A: General. 2009 , 368(1-2), 87-96

Pachon, L. D.; Rothenberg, G. Appl. Organometal. Chem. 2008 , 22 , 288-299

Pan, X.; Bao, X., Reactions over catalysts confined in carbon nanotubes. Chemical

Communications (Cambridge, United Kingdom) 2008 , (47), 6271-6281.

Peiyu, W.; Qingshan, L.; Jiangong, L. Materials Research Bulletin. 2010 , 45 , 129-

134

127

Rac, B.; Molnar, A.; Forgo, P.; Mohai, M.; Bertoti, I. J. Molec. Catal. A: Chem.

2006 , 244 , 46-57.

Rahman, I. A.; Jafarzadeh, M.; Sipaut, C. S. Cer. Int.

2008 , 35 , 1883.

Reetz, M. T.; Breinbauer, R.; Wannlnger, K.

Tetrahedron Letters, 1996, 37(26),

4499-4502

Reetz, M. T.; Helbig, W. J. Am. Chem. Soc. 1994 , 116 , 7401-7402

Reetz, M. T.; Quaiser, S. A. Angew. Chem. Int. Ed. 1995 , 34 , 2240-2241

Ren, L. Z.; Meng, L. J. EXPRESS Polymer Letters 2008, 2, (4), 251-255.

Risse, T.; Shaikhutdinov, S.; Nilius, N.; Sterrer, M.; Freund, H. J. Acc. Chem.Res

.

2008 , 41 , 949

Roucoux, A. Top. Organomet. Chem.

2005 , 16 , 261-279.

Roucoux, A.; Nowicki, A.; Philippot, K. Nanopart. Catal.

2008 , 349-388.

Roucoux, A.; Philippot, K. Handbook of Homogeneous Hydrogenation 2007 , 1 ,

217-256.

Roucoux, A.; Schulz, J.; Patin, H. Chem. Rev . 2002 , 102 , 3757-3778.

Sato-Berr´u, R.; Red´on, R.; V´azquez-Olmos, A.; Saniger, J.; J. Raman Spectrosc.

2009 , 40 , 376-380

Satoshi, H.; Yindee, S.; Atsushi, O.; Chiemi, T.; Xiuqin, C.; Shaoming, Y.;

Hideki, S.; Masahiko, A.; Nick, S. Journal of Catalysis. 2012 , 289 , 266-271

Scaiano, J. C.; Aliaga, C.; Maguire, S.; Wang, D. J. Phys. Chem. B. 2006 , 110 ,

12856- 12859 Science 2013 , 266 , 188-193

Silvio, C.; Paul, M.; Peter, J.; Ken, G.; Christopher, K. J.; Graham, H. J. Phys.

Chem. Chem. Phys. 2003 , 5(6), 1329

Somorjai, G. A. ACS Symp. Ser.

2005 , 890 , 210-219.

Somorjai, G. A. Appl. Surf. Sci.

1997 , 121/122 , 1-19.

Somorjai, G. A.; Contreras, A. M.; Montano, M.; Rioux, R. M. Proceedings of the

National Academy of Sciences of the United States of America. 2006, 103, (28),

10577-10583.

128

Somorjai, G. A.; Kliewer, C. J. Reaction Kinetics and Catalysis Letters 2009, 96,

(2), 191-208.

Somorjai, G. A.; Rioux, R. M. Catal. Today 2005 , 100 , 201-215.

Somorjai, G. A.; York, R. L.; Butcher, D.; Park, J. Y. Physical Chemistry

Chemical Physics 2007, 9, (27), 3500-3513.

Stein, A.; Adv. Mater , 2003 , 15 , 763

Sullivan, J. A.; Flanagan, K. A.; Hain, H. Catalysis Today 2009, 145, (1-2), 108-

113.

Sulman, E.; Lakina, N.; Sulman, M.; Ankudinova, T.; Matveeva, V.; Sidorov, A.;

Sidorov, S., Stud. Surf. Sci. Catal.

2000 , 130B , 1787

Sun, K.; Liu, J.; Nag, N. K.; Browning, N. D. Rec. Res. Dev. Mat. Sci.

2003 , 4 ,

521-540.

Susumu, T.; Masatake, H.; Tetsuhiko, K.; Atsushi, U.; Yoshiko, N. Stud. Surf. Sci.

Catal. 1991 , 63, 695-704

Suzuki, K.; Yumura, T.; Mizuguchi, M.; Tanaka, Y.; Chen, C.; Akashi, M .

J. Appl.

Polym. Sci . 2000 , 77 , 2678

Suzuki, S.; Suzuki, T.; Tomita, Y.; Hirano, M.; Okazaki, K.; Kuwabata, S.;

Torimoto, T.; Cryst. Eng. Comm . 2012 , 14 , 4922-4926

Takasu, Y.; Itaya, H.; Kawaguchi, T.; Sugimoto, W.; Murakami, Y. Stud. Surf. Sci.

Catal.

2003 , 145 , 279

Tamami, B.; Allahyari, H.; Ghasemi, S.; Farjadian, F.; J. Organometallic.Chem

.

2011 , 696 , 594-599

Tamami, B.; Allahyari, H.; Ghasemi, S.; Farjadian, F.; Sohelia, G.; Iranian

Polymer Journal, 2011, 20, 699-712

Tan, C. K.; Newberry, V.; Webb, T. R.; McAuliffe, C. A. J. Chem.Soc., Dalton

Trans. 1987 , 6 , 1299

Tano, T.; Esumi, K.; Meguro, K. J. Colloid Interface. Sci. 1989 , 133 , 530-533

Teranishi, T.; Hosoe, M.; Miyake, M. Adv. Mater. 1997 , 9 , 65

Teranishi, T.; Miyake, M. Chem. Mat.

1998 , 10 , 594

129

Thathagar, M. B.; Beckers, J.; Rothenberg, G. Journal of the American

Chemical.Society.

2002 , 124(40), 11858-11859

Thibault-Starzyk, F.; Daturi, M.; Bazin, P.; Marie, O. Nanopart. Catal.

2008 , 505-

528.

Thomas, J. M.; Johnson, B. F. G.; Raja, R.; Sankar, G.; Midgley, P. A. Acc. Chem.

Res.

2003 , 36 , 20-30.

Thomas, J. M.; Raja, R. Topics in Catalysis 2006, 40, (1-4), 3-17.

Torigoe, K.; Esumi, K. Langmuir. 1992 , 8 , 59-63

Toshima, N.; Ohtaki, M.; Teranishi, T. React. Polym. 1991 , 15 , 135

Toshima, N.; Takahashi, T.; Hirai, H. Chem. Lett. 1986 , 35-38

Tsvelikhovsky, D.; Popov, I.; Gutkin, V.; Rozin, A.; Shvartsman, A.; Blum, J.

European Journal of Organic Chemistry 2009 , (1), 98-102.

Turkevich, J.; Stevenson, P. C.; Hillier, J. Discuss. Faraday Soc. 1951 , 55-75

Ungureanu, S.; Deleuze, H.; Babot, O.; Achard, M.; Sanchez, C.; Popa, M. I.;

Backov, R. Applied Catalysis, A: General. 2010, 390(1-2), 51-58

Wallow, T. I.; Novak, B. M. J. Org. Chem.

1994 , 59 , 5034

Wan, Y.; Wang, H.Y.; Zhao, Q.; Klingstedt, M.; Terasaki, O.; Zhao, D.

J.Am.Chem. Soc . 2009, 131, 4541-4550

Wang, D.; Xie, T.; Li, Y. Nano Res.

2009 , 2 , 30-46.

Wang, Q.; Ostafin, A. E. Encyclopedia of Nanoscience and Nanotechnology 2004 ,

5 , 475-503.

White, R. J.; Luque, R.; Budarin, V. L.; Clark, J. H.; Macquarrie, D. J. Chem. Soc.

Rev. 2009 , 38 , 481-494

Wood, M. A. J. Royal Soc., Interf. / Royal Soc.

2007 , 4 , 1-17.

Wu, L.; Ling, J.; Wu, Z.; Adv. Synth. Catal. 2011 , 353 , 1452-1456

Xiaoli, J.; Bo, H.; Jigen, M.; Junhua. Y.; Qixian, Z.; Li, N. Carbon , 2012 , 50 ,

3083-3091

Yadav, G. D. Catal. Surv. Asia 2005 , 9 , 117-137.

130

Yonezawa, Y.; Sato, T.; Ohno, M.; Hada, H. J. Chem. Soc., Faraday Trans.

1987 , 83 , 1559-1567

Yu, W.; Liu, H. Chem. Mater. 1998 , 10 , 1205

Yu, W.; Liu, M.; Liu, H.; Zheng, J. J. Colloid Interface Sci. 1999 , 210 , 218

Zhang, M.; Zhang, W.

Journal of Physical Chemistry C 2008, 112, (16), 6245-

6252.

Zhao, M.; Crooks, R. M. Adv. Mater. 1999 , 11 , 217

Zhao, M.; Crooks, R. M. Angew. Chem., Int. Ed. Engl. 1999 , 38 , 364

Zhao, M.; Sun, L.; Crooks, R. M. Polym. Prepr . 1999 , 40 , 400

Zhi, J.; Song, D.; Li, Z.; Lei, X.; Hu, A. Chem. Commun.

2011 , 47 , 10707–10709

131